You are on page 1of 260

Optimal Design of Laminated Structures

with Local Reinforcements

Benjamin Schlpfer

DISS. ETH NO. 20894

Optimal Design of
Laminated Structures with
Local Reinforcements

A dissertation submitted to

ETH Z URICH
for the degree of

Doctor of Sciences
presented by

B ENJAMIN G EORG S CHLPFER


MSc ETH Masch.-Ing.
born September 13, 1983
citizen of Wald (AR)
accepted on recommendation of
Prof. Dr. E. Mazza, examiner
Prof. Dr. W. Becker, co-examiner
Dr. G. Kress, co-examiner
2013

Abstract
Composite materials, such as fiber-reinforced plastics, have outstanding
mass-specific mechanical properties along the fiber reinforced direction, wherefore they are increasingly used for light-weight applications. They are highly
anisotropic due to the much lower stiffness and strength values in directions
perpendicular to the fibers. Consequently, they are typically used as layers
building a multidirectional laminate. The design parameters of laminates include the fiber orientations, the laminate thickness or the number of layers,
respectively, their stacking position and the choice of the material. They can be
used to tailor the laminate properties to local stiffness and strength requirements under defined loadcases. The internal load distribution of a structure
depends on the problem domain geometry and boundary conditions and is generally inhomogeneous and direction-dependent. The search for best structural
performance, particular regarding lightweight applications, must include an
adaption of the local laminate properties to the local load distributions. Aligning the local anisotropies of the material elements according to the load directions improves the material utilization.
This dissertation investigates the optimal design of laminated composites
with local reinforcements. The work introduces a new sensitivity- and finiteelement-based method, called Ghost Layer Method, for finding structural designs with much improved mechanical performance. Sensitivities of minimum
compliance, natural frequencies, displacement, stability and strength are derived analytically with respect to the layer thicknesses of each finite element
which allows an efficient evaluation without numerical errors. They provide
information on how and where to adapt the laminate properties in order to
change the objective most effectively in terms of additional mass. The underlying parameterization concept predefines layers with given fiber orientations and stacking positions which guarantees the obtained solutions to be
manufacturing-friendly. A ghost layer is defined to have material properties
and a fiber orientation but its thickness is specified to zero wherefore it does
not contribute to the structural performance. However, its sensitivity values
are non-zeros which allows generating information on the structural design by
adding a set of these layers at the interfaces of the regular layers. Local reinforcements are generated by assigning a real thickness to parts of the ghost
layers according to the sensitivities.
For verification, a set of different structural design problems of academic
nature, which can be proved to be optimal, are directly compared to the corresponding solutions obtained with the presented method. The verification

II

indicates that the method finds optimal design solutions for convex problems
and it may be expected that the method steps towards an optimal solution
starting from an initial design in search space. A set of laminate design solutions is presented regarding eigenfrequencies, nodal displacements, stability
and strength. A panel, whose natural frequencies are shifted by the application
of local reinforcements in order to create a resonance-free band, and a locally
reinforced notched plate with increased strength performance are physically
realized using a sequential curing process. There, the basic laminate and the
reinforcements are cured and bonded in two separate temperature cycles. The
laminate quality is good and the complex reinforcement contours can be patterned exactly. According to the results of the validation, the eigenmodes and
eigenfrequencies of the panel, which are measured with a scanning vibrometer, match the simulation results very well and the resonance-free band is confirmed. The strength increase of the notched plate, which is experimentally determined with digital image correlation and acoustic emission measurements,
is less high than predicted but still high with respect to the additional material needed. The presented design process allows to efficiently generate locally
reinforced laminate designs having an outstanding structural performance.

Zusammenfassung
Verbundmaterialien, wie zum Beispiel faserverstrkte Kunststoffe, haben
hervorragende masse-spezifische mechanische Eigenschaften in Richtung der
Verstrkungsfasern, weshalb sie besonders im Leichtbau verwendet werden.
Aufgrund der signifikant schlechteren Steifigkeits- und Festigkeitseigenschafen quer zur Faserrichtung sind sie stark anisotrop. Deshalb werden sie typischerweise als Lagen in einem multidirektionalen Verbund verwendet. Die
Design-Parameter von Laminaten sind die Faserorientierung, die Laminatdicke respektive die Anzahl Lagen, die Schichtreihenfolge sowie das Material
selbst. Sie knnen verwendet werden, um die Laminateigenschaften auf die
lokalen Steifigkeits- und Festigkeitsanforderungen, bezglich definierter Lasten, anzupassen. Die interne Lastverteilung einer Struktur ist abhngig von
der Bauteilgeometrie sowie von den Randbedingungen und ist im Allgemeinen orts- und richtungsabhngig. Die Suche nach dem besten Strukturdesign,
speziell bei Leichtbauanwendungen, fordert eine Anpassung der lokalen Laminateigenschaften an die lokale Lastverteilung. Die Ausrichtung der lokalen
Anisotropie bezglich der Lastrichtung verbessert die Materialausnutzung.
Diese Dissertation untersucht die optimale Gestaltung von laminierten Faserverbundwerkstoffen mit lokalen Verstrkungen. Die Arbeit stellt eine neue,
Sensitivitts- und Finite-Elemente-basierte Methode, genannt Ghost Layer
Methode, fr das Finden von Strukturlsungen mit verbesserten mechanischen Eigenschaften vor. Sensitivitten bezglich minimaler Nachgiebigkeit,
Eigenfrequenzen, Verschiebung, Stabilitt und Festigkeit werden analytisch
bezglich der Lagendicken in jedem Finiten Element hergeleitet, wodurch eine
effiziente Auswertung ohne numerische Fehler mglich wird. Sie bieten Informationen, wie und wo die Laminateigenschaften angepasst werden mssen,
um mit geringster zustzlicher Masse mglichst grossen Einfluss auf das Optimierungsziel zu haben. Das zugrunde liegende Parametrisierungskonzept definiert Lagen mit vorgegebener Faserorientierung und Schichtreihenfolge, wodurch sichergestellt wird, dass fertigungsnahe Lsungen entstehen. Ein Ghost
Layer hat definierte Materialeigenschaften und Faserorientierung, besitzt jedoch keine Dicke, womit er keinen Beitrag zum Strukturverhalten leistet. Seine Sensitivitten verschwinden jedoch nicht. Durch das Platzieren dieser Lagen an den Grenzflchen der realen Schichten knnen zustzliche Informationen ber das Strukturdesign generiert werden. Die lokalen Verstrkungen entstehen, indem, unter Bercksichtigung der Sensitivitten, Teilen dieser Ghost
Layer eine reale Dicke zugewiesen werden.

IV

Eine Auswahl an Strukturproblemen akademischer Natur, fr welche Optimalitt nachgewiesen werden kann, wird mit den entsprechenden Lsungen
der prsentierten Methode verglichen. Die Verifikation zeigt, dass die Methode
optimale Lsungen von konvexen Strukturproblemen findet. Fr komplexere
Probleme kann angenommen werden, dass die Methode, ausgehend von einer
Startlsung, in Richtung der optimalen Lsung schreitet. Eine Auswahl von
Laminatlsungen fr Strukturprobleme bezglich Eigenfrequenz, Knotenverschiebungen, Stabilitt und Festigkeit werden prsentiert. Ein Panel, dessen
Eigenfrequenzen durch die Anwendung von lokalen Verstrkungen verschoben
werden um ein Resonanz-freies Band zu generieren, sowie eine lokal verstrkte, gelochte Platte mit gesteigerten Festigkeitseigenschaften werden in einem
sequentiellen Aushrtungsprozesses realisiert. Dabei werden das Grundlaminat und die Verstrkungen in zwei separaten Temperaturzyklen ausgehrtet
und miteinander verbunden. Die Laminatqualitt ist gut und die Konturen der
lokalen Verstrkungen knnen exakt nachgebildet werden. Die Eigenmoden
und die Eigenfrequenzen des Panels, welche mit einem Scanning-Vibrometer
experimentell ermittelt werden, stimmen sehr genau mit den Resultaten aus
der Simulation berein und die Existenz eines Resonanz-freien Bands wird
besttigt. Die Festigkeitssteigerung der gelochten Platte, welche mit Hilfe von
Bildkorrelation und akustischen Methoden gemessen wird, ist weniger hoch
als erwartet aber trotzdem in einem guten Bereich, bezogen auf das zustzliche Gewicht. Die vorgestellte Designmethode ermglicht eine effiziente Generierung von lokal verstrkten Laminatlsungen mit hervorragenden strukturellen Eigenschaften.

Acknowledgements
This thesis has been carried out at the Institute for Mechanical Systems at
ETH Zrich. The work has been supported by the Commission for Technology and Innovation (CTI-project 10279.1) and the the Swiss National Science
Foundation (SNSF-project 200021-131921).
Numerous people have contributed in various and important ways to this thesis. My special thanks go to:

The Swiss Federation and its tax payers for giving me the great opportunity to enjoy this outstanding education.
Prof. Edoardo Mazza for being my supervisor and Prof. Paolo Ermanni
for giving me the opportunity to carry out this work at ETH.
Dr. Gerald Kress for being my doctoral advisor, for help, corrections and
support.
Prof. Wilfried Becker for being my co-examiner.
The workmates Florian Bachmann, Oliver Hfner, Alex Hasse, David
Keller and Thomas Mayer for any kind of support and for making me
laugh.
The students Benedikt Rentsch, Benjamin Schneuwly, Veit Gabatuler
and Marco Marinelli for their work contributing to this thesis.
Dr. Andreas Brunner for carrying out the acoustic measurements.
All my friends for letting me enjoy the time beyond the work.
My parents and my family for their lifelong support.

December 2012, Benjamin Schlpfer

VI

Contents
Notation
1 Introduction
1.1 Laminated Composites . . . . . .
1.1.1 History . . . . . . . . . . . .
1.1.2 Material characteristics .
1.2 State-of-the-art . . . . . . . . . . .
1.2.1 Laminate Design Process .
1.2.2 Laminate Optimization . .
1.2.3 Commercial Solutions . . .
1.3 Environment, Motivation & Goal
1.3.1 Initial Situation . . . . . .
1.3.2 Project Funding . . . . . .
1.3.3 Motivation and Goal . . . .
1.4 Thesis Outline . . . . . . . . . . .

XI

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

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

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

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

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

1
1
1
3
5
5
10
16
17
17
18
18
20

2 Fundamentals
2.1 Governing Equations of Linear Elasticity . . . . . . . . .
2.2 Classical Laminate Theory . . . . . . . . . . . . . . . . . .
2.2.1 Plane-Stress State and Stiffness Transformation
2.2.2 Stiffness Homogenization . . . . . . . . . . . . . .
2.3 Laminate Failure Criteria . . . . . . . . . . . . . . . . . .
2.3.1 Tsai-Hill Criterion . . . . . . . . . . . . . . . . . . .
2.3.2 Tsai-Wu Criterion . . . . . . . . . . . . . . . . . . .
2.4 Finite Element Method . . . . . . . . . . . . . . . . . . . .
2.4.1 Equation of Motion . . . . . . . . . . . . . . . . . .
2.4.2 Stability Problems . . . . . . . . . . . . . . . . . . .
2.4.3 Layered Shell Elements . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

21
21
26
27
29
32
33
34
35
36
39
41

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

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

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

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

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

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

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

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

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

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

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

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

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

VIII

CONTENTS

3 Sensitivities
3.1 Finite Element Based Sensitivities . . . . . . .
3.1.1 Stiffness Matrix Sensitivities . . . . . .
3.1.2 Mass Matrix Sensitivities . . . . . . . .
3.1.3 Stress Stiffness Matrix Sensitivities . .
3.1.4 Displacement Sensitivities . . . . . . . .
3.2 Eigenfrequency Sensitivities . . . . . . . . . . .
3.3 Minimal Compliance Sensitivities . . . . . . . .
3.4 Sensitivities of Nodal Displacement . . . . . .
3.5 Buckling Load Factor Sensitivities . . . . . . .
3.5.1 Simplifications . . . . . . . . . . . . . . .
3.6 Strength Sensitivities . . . . . . . . . . . . . . .
3.6.1 Pseudo Strength Function Sensitivities
3.6.2 Failure Criteria . . . . . . . . . . . . . .
3.7 Hybrid Evaluation . . . . . . . . . . . . . . . . .
3.8 Smoothing . . . . . . . . . . . . . . . . . . . . . .

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

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

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

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

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

49
51
51
56
57
58
61
65
68
71
73
76
77
77
83
84

4 Automated Design Process


4.1 Ghost Layer Concept . . . . . . . . . . . . . . . . . . . . .
4.2 Parameterization . . . . . . . . . . . . . . . . . . . . . . . .
4.3 Search Direction . . . . . . . . . . . . . . . . . . . . . . . .
4.3.1 Search Direction as Design Support Information
4.4 Adaption Process . . . . . . . . . . . . . . . . . . . . . . . .
4.4.1 Limitations . . . . . . . . . . . . . . . . . . . . . . .
4.4.2 Comment . . . . . . . . . . . . . . . . . . . . . . . .
4.5 Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.5.1 Model Requirements . . . . . . . . . . . . . . . . .
4.5.2 Mesh Size . . . . . . . . . . . . . . . . . . . . . . . .
4.5.3 Physical Errors . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

89
89
92
93
96
98
104
105
105
105
106
109

5 Verification
5.1 Problem Formulation . . . . . . . . .
5.1.1 Finite Element Model . . . . .
5.1.2 Reference Solutions . . . . . .
5.2 Minimal Compliance . . . . . . . . . .
5.2.1 Loadcase 1: Single load . . . .
5.2.2 Loadcase 2: Distributed load .
5.3 Minimal Deflection . . . . . . . . . . .
5.4 Eigenfrequency . . . . . . . . . . . . .
5.5 Buckling Load Factor . . . . . . . . .
5.6 Strength . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

115
115
115
116
118
119
123
126
127
133
136

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

CONTENTS

IX

5.7 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6 Laminate Applications
6.1 Vibrating Panel (Eigenfrequency) . . . . . . . . . . . .
6.1.1 Maximization of the Lowest Eigenfrequency .
6.1.2 Maximization of -Weighted Eigenfrequencies
6.1.3 Eigenfrequency Splitting . . . . . . . . . . . . .
6.2 Grabber (Nodal Displacement) . . . . . . . . . . . . . .
6.3 Buckling Plate (Buckling) . . . . . . . . . . . . . . . . .
6.4 Notched Plate (Strength) . . . . . . . . . . . . . . . . .

136

.
.
.
.
.
.
.

137
137
138
142
144
149
151
155

7 Manufacturing-Related Aspects
7.1 Sequential Curing Process . . . . . . . . . . . . . . . . . . . . . .
7.2 Reinforcement Layer Manufacturing . . . . . . . . . . . . . . . .
7.3 Process Improvements . . . . . . . . . . . . . . . . . . . . . . . . .

161
162
166
171

8 Validation
8.1 Vibrating Panel . . . . . . . . . . . . . . .
8.1.1 Panel with Basic Laminate . . . .
8.1.2 Panel with Local Reinforcements
8.2 Notched Plate . . . . . . . . . . . . . . . .
8.2.1 First Ply Failure . . . . . . . . . .
8.2.2 Ultimate Failure Load . . . . . .
8.3 Discussion . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.

173
173
175
176
178
180
191
193

9 Discussion and Conclusion


9.1 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.2 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

195
195
198

A Material Properties
A.1 T300/Epoxy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
A.2 Aluminum 2014-T6 . . . . . . . . . . . . . . . . . . . . . . . . . .

i
i
ii

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

B Stiffness Matrix Sensitivities


C Beam Finite Elements
D Mode Comparison of the Vibrating Panel
D.1 Modes of the Basic Panel . . . . . . . . . . . . . . . . . . . . . . .
D.2 Modes of the Reinforced Panel . . . . . . . . . . . . . . . . . . . .
Acronyms

iii
v
vii
vii
x
xiii

Bibliography

CONTENTS

xvi

Own Publications

xxxv

Curriculum Vitae

xxxvii

Notation
A
A(x)
B
C
C i jkl
c
E
e
kl

0
F
fi
f
FI

kl
I z (x)
K
k

K
L
L

cr
M

Element area
Variable cross section area
Strain-displacement matrix
Material stiffness matrix
Material stiffness tensor
Wave propagation speed
Youngs modulus
Specific strain energy
Strain tensor
Strain vector
Strain
Membrane strain vector
Single force
Internal body force tensor
Internal body force vector
Failure index
Surface of the domain
Engineering shear strains
Area moment of inertia (along z-axis)
Global stiffness matrix
Element stiffness matrix
Plate curvatures
Shear correction factor
Global stress stiffness matrix
Differential matrix operator
Lagrangian
Eigenvalue
Buckling load factor
Global mass matrix

XII

M
m
M
Mb
Mb
(x)
N

x , y

Q
Q
q
R
R
r

S
S
s
s
s(x)
i j

T
T
t
tmat
t
t ctrl
t(x)
U
ui
u

Notation

Line moment / Moment per unit length


Element mass matrix
Lumped mass
Bending moment
Bending moment caused by a dummy unit load
Mass per unit length / Specific structural weight
Line force / Force per unit length
Angular frequency
Domain of the finite element problem
Sub-domain
Eigenvector
Ply orientation angle
Plane rotations of the normal to the middle plane
Shape function
Potential energy
Reduced stiffness matrix
Transformed reduced stiffness matrix
Distributed force
Reuter matrix
Strength ratio
Load vector
Material density
Material compliance matrix
Shear strength
Initial stress matrix
Search direction
Specific stiffness
Cauchy stress tensor
Cauchy stress vector
Surface stresses
Kinetic energy
Rotation matrix
Laminate thickness
Material thickness array
Element-layer thickness array
Thickness control array
Variable thickness
Deformation energy
Displacements
Displacement vector

Notation

u
W
X
Xc
Xt
Y
Yc
Yt

XIII

Prescribed displacements
Nodal point displacement vector
Potential of external forces
Strength in fiber direction
Compression strength in fiber direction
Tensile strength in fiber direction
Strength perpendicular to fiber direction
Compression strength perpendicular to fiber direction
Tensile strength perpendicular to fiber direction

XIV

Notation

Chapter 1

Introduction
1.1 Laminated Composites
1.1.1

History

Aviation is posing the challenge of building load carrying structures with minimal weight since the start of the 20th century. The Wright Brothers built
their first aircraft of wood and cotton fabrics. Wood is a good light weight material and therefore still used for special aircraft applications, e.g. acrobatic
flight vehicles. Between World War I and World War II, also called the Golden

Figure 1.1: Wilbur & Orville Wright with a glider, 1901 [120]

Introduction

Age of Aviation, the aircraft technology made great advancements. Due to


the more sophisticated aircraft performance requirements, wood was replaced
by metallic materials, mainly aluminum and steel. The developments were
driven by the numerous competitions such as the non-stop flight by Lindbergh
crossing the Atlantic. During World War II, the development and the production of aircrafts increased drastically. Huge advancements have been made in
the field of mass production. The economic boom after the wars effected a rapid
growth of the civil aircraft industry. Saving mass became more important since
the reduction of fuel consumption extends the operating distance. During the
Cold War, the research of light weight design was pushed by the defense- and
the space industries of the United States and the Soviet Union. Besides new
aluminum alloys, fiber reinforced composites have been introduced for specific
applications. Composite materials, especially fiber reinforced plastics (FRP),
have outstanding mass-specific mechanical properties compared to common
metallic material such as aluminum or steal wherefore they are predestined
for light-weight structures. The fuselage section and the horizontal tail of the
General Dynamics F-111 were some of the first high performance components
made of composite (in this case a boron/epoxy laminate) [72]. The first carbon
fiber reinforced plastics (CFRP) component was the fuselage of the F-5 fighter.
More composites were used for the fourth generation fighters such as the F-14,
F-15 and F-16. However, the usage of laminated composites has been mainly
restricted to military and space applications since the costs of these materials are high and the manufacturing process is more difficult if compared to
that of metallic solutions. In the 1980s, the first composite parts have been
used in commercial aircrafts, e.g. in the Boeing 767, the Antonov 124 and
the Airbus 310. Besides the research in material science, the development of
computer-based design processes has been pushed successfully. The formulation of the Finite Element Method (FEM) enabled the design of structures with
a higher degree of complexity which provides the potential of finding solutions
with better light weight properties. The FEM plays a decisive role for understanding the complex behavior of anisotropic laminates and is essential for the
design of light weight structures with composite materials.
The usage of laminated composite materials is today already state-of-the-art
in the sports sector, especially in areas where low weight is essential. With
the exception of the engine, gear components and wheel carriers, Formula 1
racing cars are completely made of CFRP materials. The so-called monocoque,
which protects the driver in case of a crash, is a single piece laminated carbon
part. Most of the parts of racing sailing boats (e.g. the Americas Cup Class)
are made of composites. Moreover, composite golf shafts, hockey sticks, tennis
rackets or bicycle frames are prevalent in professional and amateur sport.

1.1 Laminated Composites

With the new understanding of the global warming and the increasing price
of fossil fuels, the reduction of fuel consumption becomes interesting not only
for the aerospace industry but for the entire transportation sector. Besides of
the efficiency of the engine system, the weight of a vehicle has a strong impact on the fuel consumption. Today, the higher costs of FRP materials can
be compensated with lower operating costs. The first generation of large civil
aircrafts, where the primary structure including fuselage and wing are made
of composite materials, has been put into operation recently. The reduction
of the fuel consumption, which is effected by more efficient engines, improved
aerodynamic design and the replacement of aluminum with CFRP is in the
range of 20%. Currently, a huge research effort is done by the automotive industry which is driven by the increasing fuel costs, strict regulations by the
governments and a general demand for low consumption vehicles. While some
high-end cars are almost entirely made of laminated composites, the industry
tries to enhance the technology to mass production in order to provide light
cars at affordable prices.

1.1.2

Material characteristics

Composite materials are a combination of at least two different sub-materials.


The combination aims for a third material having superior properties by taking advantage of the properties of its sub-materials. On a macroscopic scale,
the composite material can be considered as homogeneous which simplifies the
design and analysis significantly. The composite may have material properties
that cannot be achieved by the single components. Since not all properties can
be improved simultaneously, the composites have to be designed considering
the specific requirements of an application.
The most important class of composites are the Fiber Reinforced
Plastics (FRP) which consist of fiber material embedded in matrix material.
The superior material properties of FRP are mainly arising from the fiber material. Typical fiber materials are carbon, glass, boron or aramid. The benefit of fiber materials is the low number of voids due to their small diameters
which restrains crack growing and therefore increases the strength [47]. Moreover, the mentioned fiber materials have low densities compared to common
metallic materials (e.g. carbon 1.8 g/cm3 ) which results in outstanding massspecific mechanical properties. However, the fibers themselves are not able to
build a continuous material wherefore they are embedded in a matrix system
(see Figure 1.2(a)) which are preferably duroplastic or thermoplastic materials.
Widely spread matrix materials are epoxy resins which are easy to process and
have acceptable mechanical properties with low specific weight ( 1.2 g/cm3 ).

Introduction

In order to simplify the handling in the manufacturing process, FRP are usually prefabricated. Fiber rowings are joined to laminar plies, the so-called laminae. The fibers are processed to textiles with two or multiple fiber directions
or to unidirectional plies with a single principal fiber orientation. In so-called
PREPREGs, the fiber plies are already pre-impregnated with the polymer resin.
This ensures an ideal mixing ratio of the components and a simple handling of
the fabrics which results in a high quality of the final laminated composite.

(a) Matrix-embedded unidirectional fibers

(b) Laminated composite

Figure 1.2: Laminated composites


The mechanical properties of such FRP plies are highly anisotropic. While
the mass-specific properties in fiber direction may be one magnitude higher
than for metallic materials, the material properties perpendicular to the fiber
orientation, which are mainly governed by the properties of the matrix material, are comparatively low. Therefore, structures are built by stacking several
plies with different orientation angles which are then called laminated composites (see Figure 1.2(b)). Basically, the orientation angles can be chosen in
a way so that the overall mechanical behavior becomes nearly isotropic (also
called quasi-isotropic). This is achieved by distributing the fiber orientations
regularly in all directions. A big advantage of laminated composites is the possibility to specifically design their structural response to different loadings and
requirements. In contrast to isotropic materials for which the amount of material, in particular the thickness, is the only design parameter, the behavior of
composites can be designed by varying the orientation angles, the number of
plies, the stacking sequence or in special cases the fiber volume content. Material can be added easily to highly loaded regions and the fibers can be oriented
dependent on the principal directions of the loads. Furthermore, the layered
design enables to build laminates of different ply materials. Due to the behavior of anisotropic materials, the layered design method and the resulting
large number of design variables, the design process of laminated composites

1.2 State-of-the-art

becomes complex and time consuming. The experience and the intuition of
structural engineers may be pushed to the limit and, excepting structures with
a low degree of complexity, e.g. plates or cylinders, finding solutions with good
light weight properties is hard to be accomplished manually. The application
of computer-based design methods is indispensable aiming for high quality designs in a reasonable time frame.

1.2 State-of-the-art
1.2.1

Laminate Design Process

In general, laminated composite structures are designed rather conservatively


and the weight savings primarily arise from the good mechanical properties
of the material. Tsai comments the current design methods in his book saying that "fortunately, the polymer-matrix composites are so strong that they
have been reliable and competitive in spite of the less-than-perfect design practice" [167]. Taking the aerospace industry as an example, the guidelines for
the design are very strict since failure of a structural part would have disastrous consequences for which nobody would take responsibility. Basically, the
orientation angles of the unidirectional composite layers are restricted to 0 ,
+45 , 45 and 90 . Additionally, the percentage of the fibers in every of the
four directions must lie between 10% and 50% and the layup is restricted to
symmetric and balanced laminates. This ensures a minimal stiffness and/or
strength to exist in all directions. Consequently, the most extreme anisotropic
laminate possible is for example achieved with 50% fibers in 0 , 40% in 45
and 10% in 90 direction. The overall stiffness of such a laminate is slightly
anisotropic.
Figure 1.3 shows a polar plot of the tensile stiffness of a standard
T300/Epoxy laminate (material properties see Appendix Table A.2) for different
laminate configurations. The values are normalized with the stiffness of aluminum in order to enable a comparison with a common metallic solution. The
stiffness of aluminum is indicated with the black dotted line and maps onto the
unit circle. The red curve indicates the stiffness of a unidirectional laminate
which demonstrates the most anisotropic configuration possible. A maximal
value of 130 GPa is reached in fiber direction (0 ) which is 1.86 times the stiffness of aluminum. Accordingly, the minimum stiffness of 9 GPa is reached
perpendicular to the fiber direction at 90 . The blue circle indicates the stiffness of a quasi-isotropic laminate composed of T300/Epoxy which has a layup
of (0/45/-45/90)S . Surprisingly, the stiffness is remarkably lower compared to
a standard aluminum solution. The green curve shows the mentioned most

Introduction

90

2
60

120
1.5

150

30

180

210

330
unidirectional
quasiisotropic
240

300
270

aerospace max
aluminum (reference)

Figure 1.3: Tensile stiffness T300/Epoxy S1 1()/S11,alu

extreme anisotropic laminate for aerospace applications. The stiffness exceeds


the stiffness of aluminum only slightly in a range of 20 . Overall, it may
seem that the aluminum can compete without problems with a T300/Epoxy
laminate. However, the lower density of the composite material has not been
taken into account yet. Figure 1.4 illustrates the same curves but their values
are normalized with respect to the densities. The values can be understood as
a mass specific stiffness. Considering again the blue curve, it becomes clear
that the quasi-isotropic laminate surpasses the aluminum properties only because of the lower density. Consequently, the thickness of a quasi-isotropic
laminate exceeds the thickness of an aluminum solution with equivalent membrane stiffness. The higher thickness increases its bending stiffness over that
of the thinner aluminum sheet and with it its buckling strength. This adds
to the lightweight potential of the composite. The composite solution will be
lighter but its price will be significantly higher due to increased amount of
material needed and the more complex manufacturing process.
The design process of composites becomes even more complex if strength
properties are taken into account. While the strength of an isotropic material

1.2 State-of-the-art

90
3
60

120

2
30

150
1

180

210

330
unidirectional
quasiisotropic
240

300
270

aerospace max
aluminum (reference)

Figure 1.4: Tensile stiffness T300/Epoxy normalized

S 11 ()

 

/ S11

alu

can be characterized with only one or two1 parameters, at least five parameters
are needed for a reasonable evaluation of the composite strength. The identification of these parameters is time-consuming and the statistical spread may be
wide. The aerospace industry mitigates this problem by using a strain-based
design criterion. They require that the maximal strain occurring anywhere
in the structure and in any direction may not exceed a given threshold which
is usually between 0.3% and 0.4%. Instead of five strength parameters, only
the minimal failure strain is needed which simplifies the design process significantly. This value is usually reached in the fiber-perpendicular direction.
Thus, matrix cracking can be avoided which is the primary failure event excluding some exceptions, e.g. uniaxial laminates under tensile load. Along the
fibers, this approach is rather conservative since a structure is designed well
when the margin against failure is homogeneous in all directions. However,
it is assumed here that the laminate fails at 0.3% strain, independent of load
1 The strength properties of isotropic materials are often characterized by two parameters, e.g.
the tension and compression yield strength. Common yield criteria such as Von Mises or Tresca
consider only one equivalent stress parameter since the tension and compression yield strength
are usually almost identical for metallic materials.

Introduction

90

6
60

120

4
30

150
2

180

210

330
unidirectional
quasiisotropic
240

300
270

aerospace max
aluminum (reference)

Figure 1.5: Tensile strength T300/Epoxy Tsai-Wu R t ()/R0.2,alu

state or failure mode. The failure strain in fiber direction may be up to 1.8%.
There exist a few failure criteria that try to predict a damage event with more
sophisticated models. They basically require more knowledge of the considered
material in terms of material strength characterization. However, there exists
no standard criterion that should be applied preferably. It is state-of-the-art
in many companies to employ several criteria and take the results of the most
conservative.
Figure 1.5 shows a polar plot of the aluminum-normalized tensile strength.
In order to be able to express the strength of a laminate, a failure criterion
must be chosen. For this comparative study, the widely used criterion of
Tsai-Wu [169] is utilized. The yield strength R0.2,alu of an aluminum alloy
(Aluminum 2014-T6, see Appendix Table A.2) which is common for aerospace
applications has been used as reference again. The anisotropy of composite
materials becomes even more significant considering the strength properties.
For a unidirectional laminate, the tensile strength in fiber direction may be up
to two orders of magnitude higher then perpendicular to the fiber direction.
Compared to aluminum, the tensile strength of T300/Epoxy is more than 5
times higher and if normalized with density (see Figure 1.6) almost 10 times

1.2 State-of-the-art

90

10
60

120
8
6

30

150
4
2

180

210

330
unidirectional
quasiisotropic
240

300
270

aerospace max
aluminum (reference)

Figure 1.6: Tensile strength T300/Epoxy normalized Tsai-Wu

R t ( )

 

R
/ 0.2

alu

higher. However, the strength perpendicular to the fibers is very poor. In contrast to the stiffness, the strength is not equal in all direction for a quasiisotropic design. The Tsai-Wu criterion predicts first ply failure (FPF) for every single layer. Subsequently, the critical layer varies for all directions of
the load. For a quasi-isotropic laminate under uniaxial tensile load, the layer
whose fiber direction is most likely perpendicular to the load is usually critical. Due to the alternation of the critical layer, the octahedral curve results
as a consequence of the quasi-isotropic laminate with four orientation angles
(0/45/-45/90) (see Figures 1.5 & 1.6). Also the strength of the quasi-isotropic
laminate exceeds the strength of the aluminum alloy only if taking the lower
density into account. In summary, using a standard CFRP such as T300/Epoxy,
a quasi-isotropic laminate brings only small benefits from the mechanical point
of view wherefore quasi-isotropic laminates are sometimes called black aluminum. Taking into account the higher costs and the more complex manufacturing process, the use of a quasi-isotropic standard laminate does not seem
attractive.
This comparison considers only a small selection of laminates with a standard material for a simple unidirectional loadcase. However, it shows the po-

10

Introduction

tential, the freedom but also the problems of the design with composite material. From the theoretical point of view, using a quasi-isotropic laminate
makes only sense if it is loaded uniformly in all direction. Such a stress state is
unusual even if multiple loadcases are considered. But the usage of quasiisotropic laminates is widely spread. Commonly, the stresses in one direction are significantly higher than in the perpendicular direction. Moreover,
there exist applications where the stress in the main direction is more than
one order of magnitude higher than to the perpendicular direction, e.g. wind
turbine blades or helicopter rotor blades. On the other hand, a pure uniaxial stress state, in which a uniaxial laminate could be used, can usually not
been achieved for real applications as well. Considering a structure, where the
stresses are distributed homogeneous, there exists always an optimal layup
which is dependent on the material stiffnesses and/or strengths. With the exception of plates and cylinders under uniform loads, stress states are usually
dependent on the position. In order to find an optimal design, the laminate
properties must be tailored to the specific loads in different regions. This is
achieved by different design techniques. The laminar building technique of
laminated composites enables to add material to regions with increased stiffness or strength requirements or to remove it where it is less needed. Alternatively, there exist manufacturing techniques which are able to align the fibers
to the principal load directions. In general, an optimal-designed laminate takes
advantage of the anisotropic properties of the plies and the material principal
axes are roughly aligned with the principle load directions of each point. Of
course, such a material distribution cannot be found exactly due to the material conditions and the manufacturing process. The engineers must try to design laminates that come close to the optimal solutions in terms of anisotropy
utilization but are consistent with the manufacturing processes and the provided material. The design process becomes even more complex and more time
consuming and taking advantage of computer aided methods becomes indispensable.

1.2.2 Laminate Optimization


The early publications considering laminated composite materials were focusing on the analysis of laminates. Ashton et al.[6] and Jones[72] give a basic
overview of the mechanics of laminated composites which is today formally
known as the Classical Laminate Theory (CLT) (see Section 2.2). The CLT homogenizes the material properties of the plies of which the laminate consists
to a global constitutive model. In addition to the material stiffness, the fiber
orientation angles, the ply thicknesses and the stacking sequence are taken

1.2 State-of-the-art

11

into account. This theory is today state-of-the-art and used for the analysis of
laminates but it does not directly address the problem of how to design a laminate.
Having a given loadcase, the structural response and the internal laminate
stresses can be determined easily taking advantage of the CLT. However, the
inverse problem of designing a laminate for a required structural response under the given loading conditions is not trivial at all. The large number of design
variables provides a great potential for tailoring the laminate properties to the
requirements but implies a complex engineering problem. Basically, this problem can be solved with a time consuming trial-and-error approach. The solution quality depends on the complexity of geometry and loadings as well as on
skills and experience of the designer. Better solutions in less time can usually
be found by taking advantage of computational optimization methods. These
methods search iteratively for improved designs using deterministic, heuristic or stochastic mechanisms. The automated computation calls for a physical
model which is usually parameterized with the CLT.
Optimization of Simple Laminates
The first investigations on laminate optimization concepts have been initiated
by the American aerospace industry [39, 25]. Fiber volume and orientation
angles are taken as design variables and methods are restricted to simple laminates and simple loadcases. These technical reports have not been published
officially wherefore they are still hard to find.
Aiming for laminates with minimal weight, the layer thicknesses have to
be considered since neither fiber orientations nor stacking sequence are explicitly sensitive to mass. Some approaches [146, 82, 147, 137] preassign orientation angles and vary the layer thicknesses to find a weight minimal solution considering stiffness, strength and/or buckling constraints. The methods are enhanced by optimizing thickness and fiber orientation simultaneously
[162, 177, 43, 2, 9]. Bert [16] provides a method for eigenfrequency problems.
There, not only the stiffness but also the mass becomes important.
A couple of methods in literature use a sensitivity analysis to find optimal laminate layups. Sensitivities express the impact of a design variable
change on the objective function. They are built by taking the derivatives
of the objective function with respect to the design variables. The utilization
of sensitivities may increase the performance of the optimization drastically.
Moreover, they provide information on how to improve a design manually. An
overview of sensitivities in structural optimization is given by Lund [100] and
van Keulen [172]. The application of numerically derived sensitivities for laminate design is reported in [147, 3, 73]. Analytical sensitivities of orthotropic

12

Introduction

laminated rectangular plates have been formulated by Pedersen [129]. A study


on sensitivities of buckling loads with respect to fiber orientation and layer
thickness is done by Adali and Duffy [4, 33]. Mateus et al. [110] derive analytical sensitivities of finite element stiffness and mass matrices with respect to fiber orientation and layer thickness. They are processed in stiffness, buckling [109] and eigenfrequency problems [108]. Chen et al. [26] derive analytical sensitivities for eigenproblems. A more sophisticated approach
with sensitivities for non-linear laminated plates is given by Moita et al. [115].
Johansen [70, 71] maximizes the margin of safety against failure in geometrically non-linear laminated composite structures by taking advantage of fiber
orientation angle sensitivities in a 3D finite element model.
Optimization problems with sensitivity information are of course solved
with gradient-based optimization methods. In general, the algorithms assume
the objective function to be continuous, convex and at least once differentiable.
Local optima may be found if these conditions are not given which may be
acceptable in some cases, e.g. if the improvements compared to the initial design are satisfying. Gradient-based methods can be highly efficient. However,
their performance depends on the characteristics of the objective function. Due
to the trigonometric functions included in the CLT, the search space is nonconvex wherefore the methods of mathematical programming risk to get stuck
in a local optima. The theory of lamination parameters introduced by Tsai and
Pagano [168] avoids this problem by transforming the search space into a convex space. The main problem of this approach is the back-transformation of
the obtained lamination parameters to a real laminate stack which calls for a
second optimization step [7]. Fukunaga et al. [40, 43, 42] employ lamination
parameters to find strength-optimal solutions. However, they can also be applied to eigenfrequency [49], buckling [50], and stiffness problems [38].
Another approach to tackle the non-convexity of the search space is the
application of stochastic algorithms. The most popular type of stochastic algorithms in structural optimization are the so-called Genetic Algorithms (GA)
which have been first investigated by [60]. GAs are inspired by the nature simulating the mechanisms of organic evolution. The design variables
can be understood as the genes of an individual which represents a potential solution of the optimization problem. A certain number of individuals
is pooled into a so-called population. The biological principles of mutation,
recombination and selection are applied to the individuals of a population.
Their performance is usually independent on the search space topology and
they can be used for optimization problems with discrete design variables
wherefore they are versatilely applicable. Taking the stacking sequence of
the laminate as design variable, the problem of laminate optimization has

1.2 State-of-the-art

13

discrete characteristics. GAs are predestined to solve this kind of problems


[21, 85, 91, 164, 51, 125, 157, 136, 12, 84]. Moreover, they are well-suited for
multi-objective optimization [175, 132, 124]. Also non-continuous failure criteria [118] can be handled by these algorithms. The combination of GAs and lamination parameters is discussed in [164, 92, 165]. In general, the computational
costs of GAs are high wherefore they should be applied to problems where the
algorithms of mathematical programming reach their limit. This is the case
for discrete design variables, black-box optimizations2 and multi-modal topologies. Dependent on the parameterization and the structural modal, non of
them must be the case in laminate optimization.
Locally Varying Laminates
The methods mentioned above are restricted to the optimization of simple laminates. As mentioned before, the possibility of tailoring the laminate properties
to the local loading states is one of the main benefits of the laminar building
technique. A non-uniform loading state, as it is the case in almost all applications, calls for an anisotropic and inhomogeneous laminate in order to use
the material optimally. This is achieved by either locally vary the orientation angles or the laminate thickness, in particular the layer thicknesses. The
parameterization of the locally varying laminate must be taken into account
wherefore the optimization problem becomes more complex.
A large number of investigations addressing the problem of locally varying
fiber orientation angles have been published during the last years. The majority of the methods is based on a finite element framework and aims for finding
optimal orientation angles within each finite element.
P. Pedersen [130, 131] published a couple of papers for the optimal design
of plates of anisotropic materials which can be considered as fiber-net. The
principal directions of the orthotropic material within each finite element are
taken as design variables. The method is enhanced by N.L. Pedersen [127, 128]
to eigenfrequency problems. A solution of this design method is shown in
Figure 1.7(a). The lengths of the lines indicate the amount of material in the
particular directions and the stress sign distinguishes between tension (black)
and compression.
Another approach for the design of locally varying laminates is based
on homogenization methods introduced by Bendse and Kikuchi [15, 14].
Zowe et al. [193, 13, 86] enhance this approach to laminates under the name
2 Optimization problems are called black-box problems, if there is no information on the objective function and it cannot be formulated analytically. The connection between input and output
variables or the behavior of the physical model, respectively, is unknown.

14

Introduction

Free Material Optimization (FMO). It is also based on a finite element representation. The method searches for an optimal material distribution taking advantage of direct search methods whereas the homogenized material properties
within each finite element are taken as design variables. They are assumed to
be continuous which results in virtual material properties. The obtained solutions must be transformed to a real laminate layup [61].
Lamination parameters which have been mentioned above can also be
used for finding designs with locally varying laminate properties [41]. The
method is presented for eigenfrequency [1], buckling [151, 64] and strength
problems [98, 80]. Again, the back-transformation from lamination parameters to the laminate involves a second optimization problem which has been
investigated by van Campen [170, 171]. A solution of an optimal buckling stiffness plate obtained with lamination parameters is shown in Figure 1.7(b).

(a) Optimal eigenfrequency design of a clamped


plate with cutout (Pedersen[128])

 
(b) Optimal S buckling stiffness design of a simply supported, vertically loaded plate
(Setoodeh[151])

Figure 1.7: Examples of designs with locally varying fiber orientation angles
A slightly different approach for increasing the load-carrying capacity for
plates with holes through variation of orientation angles is presented by Huang
and Hafka [62].
Another approach based on finite element mesh discretization is
called Discrete Material Optimization (DMO) introduced by Stegmann and
Lund [161, 160, 102]. The method is based on ideas from multi-phase topology optimization by Sigmund [153, 44]. The material stiffness is computed as
a sum of candidate materials wherefore the discrete problem of finding an optimal material is converted to a continuous formulation where the scaling factors
are taken as design variables. The problem can be solved by taking advan-

1.2 State-of-the-art

15

tage of sensitivities and methods of mathematical programming. In general,


a different candidate material is found for each finite element which is again
problematic from the manufacturing point of view. This problem is mitigated
by enclosing a given number of elements to a region with unique laminate and
the obtained solutions can basically be manufactured with laminate patches.
Additional publications demonstrate the merits of the method for buckling and
eigenfrequency problems [103, 101].
Basically, the manufacturing of structures with locally varying orientation
angles is not easy. Nevertheless, some research groups succeed by using
so-called fiber placement machines [69, 97, 65] which is know in literature
as variable-stiffness composite laminate design. The structures have an outstanding structural response but the process is restricted to simple structures
yet and more development is needed to bring it to a wide industrial application.
Instead of varying the fiber orientation angles, the local thickness of the
laminate or in particular its layers can be varied. A few methods find optimal
thickness distributions for the single layers [126, 81]. Assuming that the thickness may be continuous, the obtained solutions are infeasible for production.
A class of methods aiming for locally varying laminates splits the design
space into subspaces assigning a unique laminate to each of them. The subspaces of these methods have to be predefined and are assumed to be constant
during the optimization. In order to guarantee the global connectivity of the
structure, a minimal number of layers have to cover multiple regions. The
mechanism of dropping single layers within the structure is also known as
laminate blending. Liu et al. [93] split a wing structure into sections and maximize the buckling load by varying the stacking sequence. A similar approach
chosen by Kristinsdottir et al. [89] is aiming for minimal weight panels considering strength constraints. GAs are used to find blended laminates for minimal weight structures considering strength [158] or buckling [5] constraints.
Seresta et al. [150] apply a GA for the design of a wing structure.
Instead of assuming the subspaces to be constant, they can be involved
in the optimization. The shapes of the laminate layers have to be parameterized in order to be able to express them with design variables. The local
laminate layers are also known as patches or reinforcement doublers. Hansel
and Becker [53] propose an adaptive topology optimization method based on a
finite element parameterization. Material is removed layer-wise considering
local element stresses and principal stress directions. The algorithm is later
enhanced to minimize an objective function by using a GA [54]. The generation
of reinforcement doublers for laminated plates with holes using this method is
presented in [34].

16

Introduction

The concept of global plies introduced in [187, 188, 186] does not only
take materials and material orientations but also the laminate patch shapes
as design variables. The shapes are parameterized with sketches in a CADenvironment and thus are restricted to a set of predefined geometries. Sections of different laminate layups in different regions arise from the patch
overlapping. A GA optimizes patch shapes, material and orientations. Alternatively, shape contours can be parametrized with spline-curves [111, 78].
More freedom in terms of patch shapes can be achieved by reintroducing a
finite-element-based representation [45, 79]. There, the patch shapes are represented with a graph-based genotype containing the adjacencies of the finite
elements in order to ensure the connectivity of the patches.
Combined Shape and Laminate Optimization
In more sophisticated approaches, the laminate optimization is combined with
a simultaneous shape optimization of the laminated structure. The combined
optimization for structural problems with stress concentrations, e.g. in vicinities of cutouts, is investigated in several publications [11, 35, 155, 95]. A simultaneous shape and laminate optimization of stringer-stiffened composite
panels is proposed in [67].

1.2.3 Commercial Solutions


All common commercial finite element software solutions (e.g. NASTRAN,
ABAQUS, ANSYS, etc.) provide the possibility of at least a simple laminate
optimization. However, the definition of design variables and objective functions is often complicated and needs a deeper understanding of the program.
The underlying optimization algorithms are made to optimize a wide variety
of objective functions which may be not ideal for the single ones. According
to experience, the obtained solutions may be difficult to comprehend and the
work expended setting up the optimization environment tends to be enormous.
Therefore, the execution of laminate optimization in industrial environments
is still not state-of-the-art. From the authors point of view, the most sophisticated design method for laminates is provided by Optistruct of Altair Engineering. The complete theory is unknown even if some parts have been published
[36, 189]. A three-step process is used to tailor the laminate to the local requirements. In the first step, the method searches for an optimal thickness
distribution for preassigned orientation angles. Subsequently, a feasible solution is generated by converting the continuous layer thicknesses to discrete
values. Finally, a stacking-sequence optimization is performed. This pragmatic method is able to find laminates with locally varying properties that

1.3 Environment, Motivation & Goal

17

are feasible for production. The three-step optimization process contains the
problem that the stacking sequence and the interaction between the layers
are neglected. However, this is essential when aiming for optimal laminates.
Especially for problems considering strength, the stacking sequence takes a
key role. Thin-ply laminates behave more like homogeneous materials since
coupling and free edge effects become less important. Moreover, the strain energy release rate per layer is less which is advantageous regarding fracture
mechanics. This is not taken into account in the first optimization step which
may result in sub-optimal designs, especially when considering strength.

1.3 Environment, Motivation & Goal


1.3.1

Initial Situation

The basic idea of the presented work has been inspired in course of a Masters thesis [143] which has been accomplished by the author in autumn 2008
at the Centre of Structure Technologies at ETH Zrich. The thesis was done
in collaboration with RUAG Space AG. An automated preliminary design process for space structures has been developed which is based on a GA. Within
a finite element model, the fundamental natural frequencies are raised by virtually placing laminate patches. Since mass must not exceed a given threshold, it is considered as a side constraint. The reinforcement patch contours
are restricted to elliptical and rectangular shapes. On the other hand, also a
mass minimization can be performed by removing material while the fundamental natural frequencies may not drop below a given value. The algorithm
is capable of finding structural solutions with increased frequency-to-mass ratios. However, its application contains some problems due to the nature of
GAs. GAs require numerous evaluations to converge. Such an evaluation always contains a finite element run which is usually performed by a commercial
solver (in this case NASTRAN). In case of the mentioned satellite model, up to
100000 evaluations are needed for obtaining a satisfying solution. GAs can be
employed reasonably if the evaluations are parallelized. Thus, many licenses
of the commercial finite element software are needed. While the number of licenses at universities is relatively high, it is usually very low in the industrial
environments. Additionally, the solutions generated with GAs are hardly comprehensible. GAs are capable of finding exceptional solutions which cannot be
found intuitively. However, these solutions are often rejected since engineers
prefer to work with solutions whose physical behavior is comprehensible. Furthermore, solutions obtained with GAs are not reproducible. They base on
stochastic mechanisms so that the terminal solutions obtained by several runs

18

Introduction

all differ from each other. Due to these reasons, the developed design process
is hardly applicable in an industrial environment.

1.3.2 Project Funding


The problem of finding improved solutions with an automated design process
is interesting from the scientific as well as the commercial point of view. Thus,
it has been decided to follow up the development of an automated design process which is better applicable in an industrial environment. A project has
been launched with RUAG Space AG which has been supported by the Commission for Technology and Innovation (CTI)3 (CTI 10279.1: Integration of an
optimization software for composite aerospace structures into the design phase
of the development process). To eliminate the mentioned problems, a new
gradient-based method with a new parameterization concept has been defined.
Starting from an initial solution, local reinforcement doublers are generated
based on gradient information in order to increase the fundamental frequencies with minimal mass addition. The promising results called for exploitation
and an enhancement of the method to design structures considering requirements as stiffness, buckling or strength. A new project has then been launched
which has been supported by the Swiss National Science Foundation (SNSF)4
(SNSF 200021-131921: Sensitivity-based optimization of laminated composite
structures with a new parameterization concept). As additional working task,
the generated solutions have been manufactured in order to experimentally
verify the simulated results.
The sets of problems addressed by the two projects are the baseline for
the presented thesis. Since the projects are closely connected, the individual
problems cannot be clearly assigned to one or the other.

1.3.3 Motivation and Goal


Even if laminated composite structures have been employed for many years
for industrial applications, they are still designed rather conservative and the
lightweight potential is often not exploited at all. Trial-and-error design methods are time consuming and the obtained solutions are less-then-perfect. The
complex material behavior and the increased number of design variables if using composite materials makes an intuitive understanding impossible which
3 The Commission for Technology and Innovation (CTI) is a Swiss promotion agency for the
support of R&D projects, entrepreneurship and the development of start-up companies. www.kticti.ch
4 The Swiss National Science Foundation (SNSF) is the most important Swiss agency promoting
scientific research. www.snf.ch

1.3 Environment, Motivation & Goal

19

calls for automated design processes. The majority of the existing methods
have some drawbacks regarding efficiency, universal validity or realization of
the obtained designs.
This dissertation faces the challenge of tailoring the laminate properties
to the local loading states in order to employ the material in regions where
it is needed and to remove it where it is useless. Taking advantage of the
anisotropy of composite materials enables to obtain better designs in terms of
material utilization which is fundamental for aiming for mass minimal solutions. An automated design method has to be developed which has the ability to
find solutions with locally varying anisotropies. It is a major requirement that
the obtained solutions are manufacturing-friendly and can be realized directly
without further extensive processing. Due to the available manufacturing capabilities, solutions with locally varying orientation angles are not feasible.
Thus, an adequate parameterization scheme is needed which enables the generation of laminates with locally varying layers which can also be understood
as reinforcement layers. Some more demands are made on the method founded
on the lacks of existing optimization practices. The parameterization scheme
and the strategy of the automated design process have to be chosen appropriately to provide an efficient process which is potentially interesting for an
application in an industrial environment. Furthermore, solution to the mentioned problem of license shortage have to be found and the required external
commercial software must be as low as possible. The design method has to
be applicable to models with arbitrary degree of complexity. Finding weight
minimal solutions makes only sense if the mass is traded off against the mechanical performance of the structure. Criteria which are primarily used in
structural design are eigenfrequency, structural stiffness, buckling load factor
and strength. The method has to be verified on a couple of selected design
problems with known solutions.
Another major goal of the thesis is to demonstrate the potential, advantages
and problems of locally reinforced structures and the manufacturing process,
respectively. The potential of the locally reinforced designs considering the
design criteria mentioned above should be presented and discussed by means
of different simulations. Improvements of the obtained results with the automated design process compared to conventional designed solutions have to
be demonstrated. A selection of designs should be manufactured and tested
to validate the method and demonstrate the advantages of locally reinforced
laminates. Consequently, an adequate manufacturing process has to be chosen which is able to realize the obtained designs as exactly as possible. The
results may additionally help to find lacks and room for improvements of the
automated design method.

20

Introduction

1.4 Thesis Outline


Chapter 2 presents some state-of-the-art fundamentals for the further derivations within this thesis. First, the governing equations of the linear elasticity
problem are recalled which are the basis for all calculations in the presented
thesis. A short introduction into the Classical Laminate Theory, which is the
standard calculus for the design of laminated composites, is given and failure
criteria are discussed. Furthermore, the fundamentals of the Finite Element
Method are presented. Finally, a general layered shell element for laminated
composites, which is used in the design process, is introduced.
Chapter 3 focuses on the derivation of the analytical sensitivities of minimal
compliance, natural frequencies, displacement, stability and strength with respect to the thickness of the respective finite element layers. A hybrid evaluation scheme taking advantage of a commercial FEM framework is introduced
which allows laminate design of more complex models and solves the problem
of license shortage. A smoothing mechanism is introduced to reduce peaks in
the sensitivity field caused by the model discretization.
The sensitivity-based automated design process is introduced in Chapter 4. It
is based on the Ghost Layer Concept which takes advantage of predefined layers with zero thickness. Local reinforcements are generated by assigning a real
thickness to parts of the ghost layers considering a search direction, which is
a manually defined weighted sum of the sensitivities. The process is discussed
regarding limitations and physical errors.
In Chapter 5, the proposed method for the design of locally reinforced structures is verified by comparing its predictions of optimum design solutions with
those of established methods.
Chapter 6 demonstrates the application of the method to more complex laminated structures. A set of laminate design solutions is presented regarding
eigenfrequencies, nodal displacements, stability and strength. A selection of
these design solutions is physically realized using a sequential curing process
presented in Chapter 7. There, the basic structure and the local reinforcements
are cured and bonded in two separate temperature cycles.
The manufactured specimens are used for the validation described in
Chapter 8. The simulation results of a vibrating panel are compared to experimental results acquired with a scanning vibrometer. Digital image correlation
and acoustic emission measurements are employed to detect matrix cracking
in notched plates during tensile tests.
The thesis is concluded with a discussion and an outlook in Chapter 9.

Chapter 2

Fundamentals
This chapter presents fundamentals on which the method presented within
this thesis is based on. A short introduction of the Theory of Linear Elasticity is given in the first section. Based thereon, the Classical Laminate Theory is presented which is the standard calculus for the mechanical behavior of
laminated composites. Subsequently, the fundamentals of the Finite Element
Method are explained and a layered shell element is formulated. The contents
of this chapter are basically state-of-the-art in the design process. For more
detailed derivations and explanations, external literature should be consulted.

2.1 Governing Equations of Linear Elasticity


The basic problem of linear elasticity can be represented with a solid body
shown in Figure 2.1. Applying kinematic and static boundary conditions, the

Figure 2.1: Solid body with kinematic and static boundary conditions

22

Fundamentals

body is deformed in an unknown way. The main objective is to find the deformation field and the corresponding strains and stresses. It is assumed that
all displacements are small and the relation between strains and stresses is
linear. The governing equations of the linear elasticity are the equilibrium
equations, the kinematic equations and the constitutive laws. The equilibrium
equation in Cartesian coordinates is formulated with the second Newtons law
on an infinitesimal volume element, which is illustrated in Figure 2.2, in the
three dimensions x1 , x2 , x3 . Newton postulates that the time rate change of

3
x3
x1

32
23

31

13
x2

12 21

Figure 2.2: Stress components on an infinitesimal volume element


linear momentum is equal to the sum of the external forces. Using a tensor
notation in the reference system with i, j = 1,2,3 , the equilibrium equations
can be written as
i j, j + f i = u i

(2.1)

Thereby, i j denotes the Cauchy stress tensor, f i the body forces, the material density and u i the displacements. The dots denote the time derivative
wherefore u i can be understood as the accelerations. The kinematic equations
connect the displacements with the strains. The strains kl can be expressed
as a function of the displacement derivatives whereas k, l = 1,2,3.
kl =


1
u k,l + u l,k
2

(2.2)

The constitutive law or the Hookes law relating stresses to strains can be expressed with a 4 th -order tensor C i jkl which holds 81 coefficients.
i j = C i jkl kl

i, j, k, l = 1,2,3

(2.3)

The equilibrium, the kinematic equations and the constitutive law can be
combined to the equation of motion that can be solved for the unknown

2.1 Governing Equations of Linear Elasticity

23

displacements u i .


1
C i jkl u k,l + u l,k , j + f i = u i
2

(2.4)

Considering a static problem, the time dependent term on the right side must
vanish wherefore the equation simplifies to


1
C i jkl u k,l + u l,k , j + f i = 0
2

(2.5)

The governing equations can be transferred into a matrix notation in which


stresses , strains and displacements u can be formulated as arrays and
the Hookes law is formulated with a stiffness matrix C. This is achieved
by taking advantage of the stress tensor symmetry (2.6), the strain tensor
symmetry (2.7) and the path independence of the deformation energy U (2.8).
i j = ji

(2.6)

kl = lk

(2.7)

U=

1
1
C i jkl kl i j = C kl i j i j kl C i jkl = C kl i j
2
2

(2.8)

These circumstances reduce the original 81 coefficients of the Hookes law to


21 which is the most general expression within the framework of linear elasticity. Such an anisotropic material definition is called triclinic and considers
no planes of symmetry for the material properties.

C11 C12 C13 C14 C15 C16

C
12 C22 C23 C24 C25 C26

C13 C23 C33 C34 C35 C36


(2.9)
C=

C14 C24 C34 C44 C45 C46

C15 C25 C35 C45 C55 C56


C16 C26 C36 C46 C56 C66
It must be taken into account that the indexing of the tensor and the matrix
formulation is different. The relations are given in Table 2.1.
Due to the consideration of the symmetry in the Hookes law, the tensor
shear strains have to be multiplied with a factor of 2 which is equivalent to the
more common engineering shear strains kl . Stresses and strains can now be
written with 6-dimensional arrays
T

= 1 2 3 23 31 12
(2.10)
T

= 1 2 3 23 31 12
(2.11)

24

Fundamentals

Stresses
matrix
tensor
1
11
2
22
3
33
4
23
5
31
6
12

Strains
matrix
tensor
1
11
2
22
3
33
4
23 = 223
5
31 = 231
6
12 = 212

Table 2.1: Indexing of tensor and matrix formulation


and the Hookes law is given as
= C

(2.12)

The number of coefficients of the Hookes law can be reduced considering the
symmetry of material properties. A material with one plane of symmetry for,
say, z = 0, which is however an arbitrary choice, is called monoclinic and has
13 independent elastic constants.

C11
C
12

C13
Cmono =
0

0
C16

C12
C22
C23
0
0
C26

C13
C23
C33
0
0
C36

0
0
0
C44
C45
0

0
0
0
C45
C55
0

C16
C26

C36

0
C66

(2.13)

Adding a second orthogonal plane of symmetry, the material constants are reduced to 9 coefficients which represents an orthotropic material. In contrast
to the materials above, there is no coupling between shear and normal components within the principal material coordinates.

C11
C
12

C13
C ortho =
0

0
0

C12
C22
C23
0
0
0

C13
C23
C33
0
0
0

0
0
0
C44
0
0

0
0
0
0
C55
0

0
0

0
C66

(2.14)

Assuming that there exists a plane in which the material properties are
equal in all directions, the number of coefficients is reduced to 5 which is the

2.1 Governing Equations of Linear Elasticity

25

transversely-isotropic case. This is evident considering laminated composites


since unidirectional reinforced plies belong to that category.

C11 C12 C13


0
0
0
C

0
0
0
12 C11 C13

0
0
0
C13 C13 C33

(2.15)
C trans =

0
0
0
0
C44

0
0
0
0
C44
0
0
0
0
0
(C11 C22 ) /2
The simplest case is the isotropic material which is described by two independent elastic coefficients.

C11 C12 C12


0
0
0
C

0
0
0
12 C11 C12

0
0
0
C12 C12 C11

C iso =
(2.16)
0

0
0
0
0
(C11 C12 ) /2

0
0
0
0
(C11 C12 ) /2
0
0
0
0
0

C
/2
(C11
12 )
The displacement vector u contains the entries of the displacements u i in the
three coordinate directions x1 , x2 , x3 .

u1
u = u2
(2.17)
u3
These coordinates can alternatively be expressed with x, y, z. In order to express the kinematic equations (2.2) in a matrix notation, a differential matrix
operator L is defined which maps the displacements u on the strains with a
simple matrix multiplication (2.19).

/ x1
0
0
0
/ x2
0

/ x3
0
0
(2.18)
L=

0
/ x3 / x2

/ x3
0
/ x1
/ x2 / x1
0
= Lu

(2.19)

The same differential operator L can be used for the matrix notation of the
equilibrium equation (2.1).

LT + f = u

(2.20)

26

Fundamentals

The combination of the equilibrium equation (2.20), the kinematic


equation (2.19) and the constitutive law (2.12) leads to a partial differential equation for the unknown displacements u.

LT CLu + f = u

(2.21)

Analogous to the tensor notation (2.5), the matrix notation of the equation of
motion can be simplified for a static problem.
LT CLu + f = 0

(2.22)

The challenge of the problem of linear elasticity is to find a displacement


solution u, which fulfills the differential equations (2.21), or (2.22) for the static
case respectively, considering stress and displacement boundary conditions.
For simple geometries and simple loadcases, e.g. for a bar under uniaxial
stress loading, a displacement solution u can be found analytically. However,
in general no closed-form solutions are available for problems with complex domain geometries, inhomogeneous material distributions, or complex boundary
conditions. The engineers have to take advantage of numerical approximation
methods such as the Finite Element Method which is introduced in Section 2.4.

2.2 Classical Laminate Theory


This section gives a short introduction of the Classical Laminate Theory (CLT)
which is the standard theory for the analysis and the design of thin laminated
composites. There exist different conventions in literature which may vary
slightly. The chosen conventions for this thesis are based on the textbook of
Jones [72]. The CLT is based on the following assumptions which arise from
the KirchhoffLove theory of plates [99]:
The laminate is assumed to be thin compared to the other dimensions
and its thickness is constant.
The assumptions of Bernoulli, namely plane sections and no transverse
shear strains, are valid.


A plane stress state is assumed z = xz = yz = 0 .

The single layers are linear elastic and bonded perfectly.


Deformations are small.

2.2 Classical Laminate Theory

27

Due to the stacking of several layers, the material stiffness properties of a


laminate through the thickness are inhomogeneous. In order to provide a linear relation between plate deformations to plate line loads of the global laminate, the CLT performs a stiffness homogenization. This requires that the
stiffnesses, which are usually given in material principal coordinates 1,2, are
transformed to the global coordinates x, y of the problem.

2.2.1

Plane-Stress State and Stiffness Transformation

The stiffness matrix of an orthotropic layer is defined with expression (2.14).


In order to express the stiffness entries as a function of the engineering constants such as the Youngs moduli, the Poisson ratios and the shear moduli, it
is easier to formulate the so-called compliance matrix S which is the inverse
of the stiffness matrix C. The compliance matrix of an orthotropic material is
defined as
1

E21
E31
0
0
0
E 11
22
33
12
1
E32
0
0
0
E 11

E 22
33
13

23
1
E

0
0
0
E 22
E 33
11

(2.23)
S ortho =
1

0
0
0
0
0

G 23

1
0
0
0
0
0
G 31
1
0
0
0
0
0
G 12
Measured engineering constants can be inserted directly and the stiffness matrix results of the inversion. Considering a unidirectional reinforced composite
material, the material properties are transversely-isotropic. As mentioned before, such a material is defined with 5 elastic coefficients which are usually the
Youngs modulus in fiber direction E 11 , the Youngs modulus perpendicular to
the fibers E 22 , the in-plane shear modulus G 12 , the in-plane Poisson ratio 12
and the transverse shear modulus G 23 . The transverse shear Poisson ratio is
related to the variables above with
23 =

E 22
1
2G 23

(2.24)

Due to the assumption of a plane stress state, the stresses with out-of-plane
components are zero and the compliance matrix S can be reduced to a 33matrix. Consequently, the stiffness matrix C is reduced to the so-called reduced
stiffness matrix Q and the Hookes law is simplified to


1
Q 11 Q 12
0
1
2 = Q 12 Q 22
0 2
(2.25)
12
0
0
Q 66 12

28

Fundamentals

Note that the strains and stresses are expressed in material principal coordinates 1,2. However, the orientation angle of a unidirectional laminate layer
can be chosen arbitrarily wherefore the material principal directions do not coincide with the global coordinates of the problem formulation. Consider a laminate illustrated in Figure 2.3 where the material coordinates 1,2 are shifted to
the global coordinates x,y by a given angle . A relation which transforms the

Figure 2.3: Material principal and global coordinate systems


stresses and the strains from the material coordinate system to the global coordinate system is needed. This is achieved by applying the rotation matrix T.

cos2
T = sin2
sin cos

sin2
cos2
sin cos

2 sin cos
2 sin cos
cos2 sin2

(2.26)

While the material stresses are mapped directly using the matrix T (as shown
in equation (2.27)),


1
x
2 = T y
12
x y

(2.27)

it has to be taken into account that the engineering shear strains kl are 2
times the tensorial shear strains (see Table 2.1) which leads to the following
relation.

1
x
2 = T y
1
1
2 12
2 x y

(2.28)

2.2 Classical Laminate Theory

29

In order to be able to work directly with the engineering strains without a


pre-factor, Reuter [141] introduced the simple matrix R.

1 0 0
(2.29)
R = 0 1 0
0 0 2
An application of this Reuter matrix reduces the potential for mistakes to a
minimum since it is clear that only engineering stains are utilized. The engineering strains are then transformed with equation (2.30).
x y = RTR1 12

(2.30)

Contrariwise, the global strains and stresses can be mapped to the material
principal strains and stresses performing a multiplication with the inverse rotation matrix T1 . The connection between the strains and stresses in global
coordinates is based on the transformed reduced stiffness matrix Q.

x
x
y = Q y
(2.31)
x y
x y
The transformed reduced stiffness matrix Q can be derived by using
equations (2.25), (2.27) and (2.30) and is therefore a function of Q, R and T
only.
Q = T1 QRTR1 = T1 QTT

(2.32)

The transformed reduced stiffness matrix Q has to be evaluated for every laminate layer in order to be able to homogenize the material data of the entire
laminate stack. However, the evaluation is straight forward and the computational costs are low.

2.2.2

Stiffness Homogenization

The theory for thin plates assumes that the displacements of a point are composed of the mid-plane displacements u0 , v0 and a term arising from the plate
curvature. The curvature term is dependent on the position through the thickness z and the rotations of the normal to the middle plane x and y . The
chosen conventions are shown in Figure 2.4.
u = u0 + z y

(2.33)

v = v0 + z ( x )

(2.34)

30

Fundamentals

z,w
z,w
y

zy

y,v

x,u

Ny
y

z w

Taking advantage of the kinematic relations (2.2), the strains yield to

x
u,x
u0,x
y,x
y = v,y = v0,y + z x,y

(2.35)

Mx

x,u

My

Nxy

Mxy

Nxy

Nx

Mxy

Figure 2.4: Loads and kinematics of a Kirchhoff-plate

x y
  

u,y + v,x




u0,y + v0,x



0

y,y x,x




where 0 denotes the mid-plane or membrane strains and the plate curvatures. Thin plate theory and therefore also the CLT assumes that the strain
distribution is linear through the laminate thickness. However, the stiffness
is discontinuous through the thickness due to the layer-wise design technique
which results in a discontinuous stress distribution. To quantify a stress state
in a laminate, the components have to be evaluated for each layer particularly.
In order to have a load unit that includes all layers, the line loads are introduced, namely the force per unit length N and the moment per unit length M.
The force per unit length N is the integration of the stress components over the
laminate thickness t.

 t x
Nx
2
y dz
(2.36)
N = Ny =
2t
Nx y
x y

Analogously, the moment per unit length M is given by the integration of the
stress components multiplied with the stacking position z.

 t x
Mx
2
y zdz
(2.37)
M = My =
2t
Mx y
x y
Since the stiffness is constant within a single layer, the integration can be
replaced with a summation of the integrals over each layer taking advantage

2.2 Classical Laminate Theory

31

zk
zk-1

hk
hk-1

zj

t
2

hj
h2
h1

z1
z0

Figure 2.5: General layup of a laminated composite material


of the transformed reduced stiffness matrix Q derived above. Additionally, the
strains are split into the membrane and the bending components arising from
equation (2.35). Consequently, the forces and the moments per unit length can
be expressed as a function of the membrane strains 0 and the curvatures .
t

N=
=

dz =

2t

n


Qj

j =1




n

j =1

z j z j 1

t

M=
=

2t

n

j =1

zdz =

Qj

Qj

n

j =1

z j
z j 1

dz =

n

j =1

Qj

z j

z j 1


0 + z dz

 
1 2
2
+
z z j 1
2 j

Qj

z j
z j 1

zdz =

n

j =1

Qj

(2.38)
(2.39)

z j
z j 1

 

 
1 2
1 3
z j z2j 1 0 +
z j z3j 1
2
3


0 + z zdz

(2.40)
(2.41)

Considering equations (2.39) and (2.41), the matrices A, B and D can be extracted which are all of dimension 33. The A-matrix connects the membrane
strains 0 with the force per unit length N.
A=

n

j =1



Q j z j z j 1

(2.42)

and the D-matrix connects the plate curvatures with the moments per unit
length M.
D=



n
1
Q j z3j z3j 1
3 j =1

(2.43)

32

Fundamentals

The B-matrix is responsible for the coupling of membrane and bending components.


n
1
Q j z2j z2j 1
(2.44)
B=
2 j =1
Dependent on the laminate layup, some entries may become zero. In case of
a symmetric laminate, there is no coupling between bending and membrane
effects wherefore the B-matrix vanishes. These matrices build the so-called
ABD-matrix which is the main achievement of the homogenization process.
  
 
N
A B 0
(2.45)
=
M
B D
Considering a laminate stress analysis problem, the loads are given and the
stresses in each layer must be found in order to check the strength. In the first
step, the ABD-matrix is calculated and the global strains can be determined
taking advantage of its inverse. The global strains are then transferred to the
local strains within each ply employing the corresponding rotation matrices T.
Finally, the reduced stiffness matrices Q map the local strains to the local
stresses which are needed for a strength analysis.

2.3 Laminate Failure Criteria


The material strength properties of unidirectional reinforced composites are
highly anisotropic. Stress limits in fiber direction are one or two orders higher
than perpendicular to it. This makes dimensioning more complicated, since the
stress state has to be compared to the particular strength values. There exists
a large number of criteria for predicting failure events. An overview with detailed explanations and experimental verifications is given in the World Wide
Failure Exercise [58]. The criteria of Tsai-Hill [57, 166], Tsai-Wu [169, 94] and
Hoffmann [59] are well-known. This is because of their simplicity in application rather than of their reliability. Their physical basis is rather weak. However, they are formulated with a quadratic closed-form expression and thus
implemented in all common commercial finite element packages. Furthermore,
the number of needed strength parameters is low. The recent and more sophisticated criteria do not only indicate a damage event but they distinguish also
between fiber and matrix cracking, e.g. Hashin [56], Puck [135] or Cuntze [30].
They are based on physical models, but their application is more expensive
since the failure indices for the several fracture modes have to be evaluated
and distinguished. For Puck, additional strength properties are required. Another approach for dimensioning composite structures is to limit the maximal

2.3 Laminate Failure Criteria

33

allowed strains. The threshold is set to the minimal breaking strain in material principal coordinates. This is reasonable when first matrix cracking cannot
be tolerated, e.g. in aircraft industry.
Within this section, the failure criteria of Tsai-Hill and Tsai-Wu are introduced
briefly. They will later be used for the laminate strength optimization in Section 3.6.2. Their closed-form expression and their simplicity in application
make them predestined for implementation for a sensitivity analysis. The fact
that their physical basis is weak can be accepted since the developed design
method is basically used for preliminary design. Furthermore, the method is
not used to design a structure to particular loads. More important it is to show
that the strength increase between initial and optimized solution is high. Thus,
the choice of the failure criterion is not so important as long as the solutions
are evaluated equally.

2.3.1

Tsai-Hill Criterion

The Tsai-Hill criterion is based on the quadratic yield condition of Hill [57] for
metals with slightly anisotropic material properties. Tsai [166] modified the
criterion by substituting the yield strength values with the strength values of
a composite material. The failure criterion for a transverse isotropic material
under plane stress is expressed with the quadratic equation

( ) =

 2
1

1 2

X2

 2
2


12 2

(2.46)

where failure is predicted if

( ) = 1

(2.47)

Thereby, X denotes the strength in fiber direction, Y the strength perpendicular to fiber direction and S the shear strength of the laminate. The criterion
does not distinguish explicitly between tensile and compression loads. However, a comparison of stress and strength makes only sense if this is considered. The user has to make sure that the appropriate values are applied.
By introducing a strength ratio R which is multiplied with the actual stresses,
the margin until failure occurs can be determined. It can be understood as a
linear scaling factor which indicates failure for a value of 1.

(R ) = 1

(2.48)

A strength ratio of 2 implies that the current load can be doubled. Since all
stresses occur quadratically in the Tsai-Hill criterion, the failure criteria with

34

Fundamentals

strength ratio can be expressed as


R2 = 1

(2.49)

The strength ratio R for the Tsai-Hill criterion is found by taking the reciprocal
of the positive square root of the failure function .
1
R= 
+

(2.50)

The so-called failure index F I which is the reciprocal of R can be used alternatively to express the margin until failure occurs. A failure index of 2 indicates
that the current loads must be reduced by a factor 2 or the laminate thickness
must be doubled in order to prevent failure.

+
(2.51)
F I = R 1 =

2.3.2 Tsai-Wu Criterion


The similar Tsai-Wu criterion [169, 94] has also quadratic characteristics. In
contrast to the Tsai-Hill criterion, Tsai-Wu contains also some linear terms.

() = F1 1 + F2 2 + F11 21 + 2F12 1 2 + F22 22 + F66 212

(2.52)

The failure condition is the same as for the Tsai-Hill criterion (2.47). The
strength parameters F of the Tsai-Wu criterion are functions of the material
strength values for tension, compression and shear.
F1 =

Xt

F11 =

Xc

Xt Xc

F2 =

Yt

F22 =

Yc

Yt Yc

F66 =

S2

The interaction term F12 for the Tsai-Wu criterion is set to


F12 = 

F12

X t X c Yt Yt

with 1 F12
1. There is no physical meaning of the factor F12
and the
choice has to be done by the user. However, recommendations for a couple of
different materials are listed in [94]. X t and X c are the strength values in fiber
direction for tension and compression, respectively. Yt and Yc are tension and
compression strength perpendicular to the fiber direction. The shear strength

2.4 Finite Element Method

35

is labeled with S.
In order to find the strength ratio R, the failure criterion (2.52) is separated
into a quadratic part Q and a linear part L .

= Q + L

(2.53)

Q = F11 21 + 2F12 1 2 + F22 22 + F66 212

(2.54)

L = F 1 1 + F 2 2

(2.55)

Consequently, the linear part L contains stresses only linear while they occur quadratically in the quadratic part Q . This has to be considered when
introducing the strength ratio R which leads to the failure condition for the
Tsai-Wu criterion.

Q R 2 + L R = 1

(2.56)

The strength ratio R can be found by extracting the positive root of the failure
condition (2.56).

L + + 4Q + 2L
= F I 1
(2.57)
R=
2Q

2.4 Finite Element Method


Within this section, a short overview of the Finite Element Method (FEM)
is given. The content is mainly based on the textbooks of Cook et al. [29],
Reddy [139] and Schwarz [149]. The FEM is a numerical method for solving
partial differential equations. It can be applied to a wide field of physical problems such as structural analysis, heat transfer, magnetic fields, flow processes
and many more. The geometry of the physical problem is discretized by dividing the domain into smaller parts called the finite elements. A finite element
has a domain, a boundary and so-called nodes on which the degrees of freedom
are defined. While the partial differential equations can only be solved analytically for simple geometries, there is no geometric restriction using the FEM.
Moreover, there is no restriction to boundary conditions or material properties
which makes the method applicable to any continuum mechanical problems.
There exist numerous commercial finite element codes which are still enhanced
and refined today. While a deeper understanding of the method requires some
effort, the usage of FEM-codes is possible with little knowledge of the method
and the underlying problem. However, the consequences of an incorrect application "may range from embarrassing to disastrous" [29].

36

Fundamentals

Within this thesis, the further explanations are focused on the FEM for structural analysis problems. The solution of the elasticity problem demands for
the fulfillment of the fundamental equations (2.21) or (2.22) within a given
and disdomain taking into consideration the prescribed surface stresses
on the surface as well as the internally acting body forces f. As
placements u
mentioned, an exact solution of the problem can only be found for simple geometries and simple boundary conditions, e.g. a bar under uniaxial load. Generally, the FEM provides only approximate solutions. Accepting a sufficient
numerical effort, the obtained solution may come close to the true solution.

2.4.1 Equation of Motion


There are several ways to derive the linear equation system of the FEM which
can be solved for the unknown displacements u. A general approach is based
on Lagrangian mechanics [46]. It is a formalism based on Hamiltons Principle
which describes the dynamics of a system with a scalar function called the
Lagrangian L and can be understood as a generalization of the principle of
virtual displacements to dynamics of solids [139]. The equation of motion in
Lagrangian mechanics, also known as the Euler-Lagrange equation, is defined
as


L
d L
=0
(2.58)

dt x
x
where L denotes the Lagrangian and x the generalized coordinates.
Lagrangian L is defined as

The

(2.59)

L = T = T (U W)

where T denotes the kinetic energy and the potential energy. The Hamiltons
principle for an elastic body is represented with equation

 t2
t1

[T (U W)] dt =

t2
t1

Ldt = 0

(2.60)

which calls the time integral of the Lagrangian L to be stationary. The fundamental lemma of the calculus of variation shows that solving the Lagrange
equations is equivalent to finding a solution of the Hamiltons principle. However, taking advantage of the Lagrangian mechanics transforms the equation
of motion to its weak form which is needed for finite element formulation.
The potential energy is the sum of the deformation energy U and the negative of the potential of the external forces W. If the displacements u are small,

2.4 Finite Element Method

37

which is required for the linear elasticity problem, the velocities can be approx The kinetic energy can thus be
imated with displacement time derivatives u.
formulated as the integral over the domain of the density and the scalar

product or the velocities u.



1
T ud

u
(2.61)
T=
2
The deformation energy U is defined as the domain integral of the scalar product of stresses and strains .

1
T d
(2.62)
U=
2
and the potential of the external forces W is dependent on the body forces f and
. Consider that the surface stresses are integrated over
the surface stresses
the surface .


T ud
(2.63)
W = fT ud +

The Lagrangian L is formulated for the entire domain . In order to get


the FEM formulation, the domain must be discretized into smaller subdomains e , namely the finite elements. The integration over the domain
is replaced with a summation of the integrals over the sub-domains e . Additionally, local approximation functions , which are also called shape func which
tions, are defined. They map the finite element nodal displacements u,
represent also the degrees of freedom, to the continuous displacements u. The
and the accelerations u
are mapped analogously.
nodal velocities u

u T u

T u
u

T u
u

(2.64)

Considering the kinematic equation (2.19), the strains can be expressed as a

function for the nodal point displacements u.


= Bu

= Lu = LT u

(2.65)

A strain-displacement matrix B, which contains the spacial derivatives of the


local shape functions, is built by applying the differential matrix operator L to
the shape functions . The total Lagrangian for the discretized system then
takes the discrete form




 1
 1 T T
T T ud
e
B CBud
e
u
u
L=
n elm 2 e
n elm 2 e






e +
e
fT T ud
T T ud
(2.66)
+
n elm

n elm

38

Fundamentals

are taken as generalized coordiFinally, the unknown nodal displacements u


nates wherefore the Lagrange-Euler equation (2.58) is modified to


L
d L
=0
(2.67)

dt u
u
The evaluation of the Lagrangian formalism leads to the equation of motion
for the discretized linear system.






e +
e
T ud
BT CBud
n elm

n elm



n elm





de = 0
fd e

n elm

(2.68)

The equation is then rearranged so that terms with no dependence on the nodal
are brought to the right hand side.
displacements u






T
T

+
B CBd e u
d e u
n elm

n elm


n elm







de
fd e +

n elm

(2.69)

The sums on the left side represent the global stiffness matrix K and the global
mass matrix M while the right hand side represents the load vector r. With
these symbols,the basic problem of the FEM for the linear elastic case is written as
= r
+ Mu
Ku

(2.70)

vanish so that the equation


Assuming a static problem, the accelerations u
simplifies to
=r
Ku

(2.71)

which is a simple linear equation system for the unknown displacements u.


The equation of motion can also be used for the determination of the harmonic
eigenfrequencies of the structural system. There, it is assumed that the load
vector r is zero. Taking advantage of the harmonic approach, the equation
of motion can be transferred into an eigenvalue problem with the unknown
eigenvalues and eigenvectors , whereas is equal to the square of the angular frequency .
= sin ( t)
u

(2.72)

= sin ( t)
u

(2.73)

2.4 Finite Element Method

39

The combination of the equation of motion (2.70) and the harmonic approach
(2.72, 2.73) yields to the eigenvalue problem for the harmonic vibration.


K 2 M = 0
(2.74)
Considering equation (2.69), the global system matrices can be extracted
directly. The global stiffness matrix K is defined as the sum of the element
stiffness matrices



BT CBd e
(2.75)
K=
n elm

whereas the element stiffness matrices k are given as



BT CBd e
k=
e

(2.76)

Analogously, the global mass matrix M and the element mass matrices m are
defined as



M=
T d e
(2.77)
n elm

and

m=

T d e

(2.78)

The load vector r contains the internal body forces and the prescribed forces on
the surface of the structure. The shape functions distribute the loads to the
nodes.






de
r=
fd e +

(2.79)
n elm

n elm

The summations must consider the connectivity of the respective element with
the globally numbered mesh nodes.

2.4.2

Stability Problems

The stress stiffness matrix K is used to determine the buckling loads of a


structure. It can be understood as stiffness part that counteracts the conventional stiffness caused by the applied loads.
K total = K + cr K

(2.80)

40

Fundamentals

In contrast to the conventional stiffness matrix K, the stress stiffness


matrix K is dependent on the loads. At the point where the stress stiffness
exactly compensates the conventional stiffness, linear bifurcation buckling occurs. The buckling load factor cr , which expresses how much the current load
may be increased until buckling occurs, is the result of an eigenvalue problem.
The equation system with stress stiffness matrix modifies to
= cr r
[K + cr K ] u

(2.81)

Since external loads do not change at the bifurcation point and the conventional stiffness matrix is not dependent on the load, buckling displacements
may be added.
+ } = cr r
[K + cr K ] {u

(2.82)

The subtraction of these two equations yields the eigenvalue problem in which
the eigenvalue cr provides the buckling load factor. The buckling mode is
consequently represented with the eigenvector .
[K + cr K ] = 0

(2.83)

The stress stiffness matrix K results from a virtual deformation energy U


caused by the nonlinear parts of the strains. The virtual work is





1 2
2
2
u,x + v,x
+ w,x
0 + ... + u,z u,x + v,z v,x + v,z v,x zx0 d (2.84)
U =
2
where 0 denotes the initial stress state (detailed derivation given in [29]).
Analogously to the regular strain energy, the virtual work for the discretized
finite element is expressed with


1
U =
n elm 2

0
u G
e
0
T

0
s
0

where s is the initial stress matrix

x0 x y0 xz0
s = x y0 y0 yz0
xz0 yz0 z0

0
e
0 Gud
s

(2.85)

(2.86)

The matrix G represents the strain-displacement relation equivalent to the Bmatrix in Section 2.4.1. However, the two matrices G and B may be different

2.4 Finite Element Method

41

for the same element wherefore a different notation is used. The most general
element stress stiffness matrix using the Lagrangian mechanics is therefore

k =

2.4.3

s
0
0

0
s
0

0
0 G dV
s

(2.87)

Layered Shell Elements

This section contains the derivation of a layered shell finite element as it has
been used for the sensitivities in Chapter 3. The derivation is separated into
stiffness-, mass- and geometric stiffness matrix. Since the design method,
which is introduced later, is coupled with the commercial finite element software NASTRAN, it was important to have a finite element that behaves as similar as possible to the ones of NASTRAN. The formulations of commercial elements are usually not available. However, numerical studies have shown that
the results coincide satisfyingly.
Stiffness Matrix
The general formulation of the element stiffness matrix k is given in
equation (2.76). Consider a layered shell element illustrated in Figure 2.6 with
an element area A and a laminate thickness t. The volume integral can be sep-

z,w
A
y,v

t
x,u
Figure 2.6: Layered shell element

arated into an integration over the area and an integration over the thickness.
k=

  t

t
2

BT CB dt d A

(2.88)

Using the CLT, the integration over the thickness can be replaced with a summation over all n layers where z represents the interface positions. Thereby,

42

Fundamentals

the material stiffness matrix C is replaced with the transformed reduced stiffness matrix Q j of the corresponding layer. It must be considered that the stiffness matrix consists of a membrane part (m), a bending part (b) and a coupling
part (c). Analogously to the derivation of the ABD-matrix, the integration
of different parts follows different rules. Moreover, the strain-displacement
relations B are different.
The derivation of the strain-displacement matrix for membrane Bm and
bending Bb starts from the kinematic relations of a laminated plate given in
equation (2.35) which are

y =

x y

u0

+ z

y x
u0 v0
+

x
x
y
x
x
v0

(2.89)

The first part describes the membrane deformation and the membrane straindisplacement matrix Bm can be written as

Bm =

(2.90)

where denote the shape functions of the finite element which are of course
dependent on the element type. The second part of equation (2.89) covers plate
bending. The resulting strain-displacement matrix Bb is
y

x y

Bb =

(2.91)

The out-of-plane deflections w are connected to the in-plane displacement u, v

2.4 Finite Element Method

43

with the transverse shear which is

w u
 
x + z
xz

=
v w
yz
+
z y

(2.92)

Alternatively it can be expressed in terms of x and y taking advantage of the


derivatives of the kinematic relations (2.33) and (2.34). The derivatives yield
u
z
v
z

= y

(2.93)

= x

(2.94)

which leads to

w
 
+

y
x
xz

w
yz
x
y

(2.95)

With that expression, the strain-displacement matrix for transverse shear is


w x y

0
y

Bs =

(2.96)

Keep in mind that the transverse shear in equation (2.92) must vanish for thin
laminates which leads to
w
x
w
y

= y

(2.97)

= x

(2.98)

Based on the strain-displacement matrices, the element stiffness matrices can


be formulated taking advantage of the laminate integration scheme of the CLT.
The coupling stiffness matrix is generated with a combination of the membrane

44

Fundamentals

and the bending shape functions.


km =

n




BT
m Q j Bm d A

A
j =1
n  


z j z j 1






1 2
T
2
BT
Q
B
+
B
Q
B
d
A
z

z
j m
m j b
j 1
b
2 j
A
j =1

n 

1 3
3
BT
Q
B
d
A
z

z
kb =
b
j
b
j 1
3 j
A
j =1


n



S
T
Bs Q j Bs d A z j z j 1
ks =

kc =

j =1

(2.99)
(2.100)
(2.101)
(2.102)

In contrast to the membrane and bending stiffness parts, the shear stiffness
matrix ks is dependent on a transverse material shear stiffness matrix QS
with dimension 22.
QS =


G 13
0

0
G 23

(2.103)

The matrix is usually multiplied with a so-called shear correction factor


which is discussed in several publications [32, 178, 163, 174, 90, 116]. The utilization of tries to overcome the physically incorrect assumption of the firstorder shear theory, namely that the shear distribution across the thickness be
constant. The complex derivation of is dependent on the laminate layup and
the ply thicknesses. However, the stiffness matrices will later be differentiated
with respect to the thickness wherefore a dependency is unwanted. Thus, a
constant factor of = 5/6, which is the value for a homogeneous isotropic plate,
is chosen for the calculation within this thesis. Considering thin elements,
the above implementation may lead to the so-called phenomenon of shear locking. The transverse shear strains become small or even negligible, and consequently the element stiffness matrix becomes stiff which yields erroneous results [138, 8, 139]. The problem is mitigated with a reduced integration of the
transverse shear stiffness matrix [190, 63, 139]. In case of a four-noded rectangular element, the transverse shear stiffness terms are integrated with the
one-point Gauss rule instead of the two-point rule. However, the four-noded
element is still prone to shear locking if the thickness is much smaller compared to the other dimensions. The influence of the shear stiffness matrix has
to be reduced artificially. This is done with an artificial shear correction factor
that considers the ratio between the element size and its thickness [83]. The

2.4 Finite Element Method

regular shear correction factor is modified with

if t/l e

 2

t
l e
a =

 2 if t/l e <

1 + tl e

45

(2.104)

For small thickness-to-length ratios, the shear artificial correction a factor


goes towards zero. The proposed method within this thesis corresponds to that
used in NASTRAN wherefore the implemented elements should be equal. The
shear correction mechanisms of NASTRAN four-node element (CQUAD4) are
not known explicitly. However, the chosen approach here matches the results
of the CQUAD4 element quite well.
The general element stiffness matrix of a layered shell k is composed with
an addressed summation of the several parts.
k = km + k c + kb + ks

(2.105)

It must be ensured that the stiffness contributions are assigned to the corresponding degrees of freedom u, v, w, x , y indicated in equations (2.90), (2.91)
and (2.96). The result of the ordered addition is thus a quadratic matrix with
dimension of 5 times the number of nodes. For modeling in a 3D-space, a sixth
degree of freedom z for the rotation around the z-axis, also called drilling
degree of freedom, is introduced. However, this degree of freedom is not important for the determination of the sensitivities in Chapter 3 and is therefore
not discussed in detail here. Investigations on elements with drilling degree of
freedom have been made in [107, 66, 28].
Mass Matrix
The mass matrix formulation for a layered shell element is simple when using a lumped mass model where rotary inertia is neglected. It is feasible for
small deformations which is usually true for harmonic vibration problems. The
lumped mass matrix is expressed with equation


n 

jtj
T d A
(2.106)
m=
j =1

The shape functions map the mass element to its nodes.


Geometric Stiffness Matrix
The general expression of the element stress stiffness matrix is given in
equation (2.87). For a plate element, a buckling instability is caused only by

46

Fundamentals

in-plane stresses ( x0 , y0 and x y0 ) wherefore the initial stress matrix (2.86)


is reduced to


x0
s =
x y0

x y0
y0

(2.107)

The spacial derivatives of the shape functions G map the nodal displacements
to the deflection derivatives w,x , w,y

G =

(2.108)

This term covers only out-of-plane buckling which is illustrated in


Figure 2.7(a). However, a general shell element should also be able to describe
buckling in the element plane as shown in Figure 2.7(b). Imagine a bar which
is modeled with shell elements and loaded with an axial force. If the element
thickness is bigger than the with, the beam should buckle in the element plane.
MacNeal [104] covers this with an additional term where the shape functions
w

(a) Out-of-plane buckling

v
(b) In-plane buckling

Figure 2.7: Buckling modes

2.4 Finite Element Method

47

G are defined as
u

G =

and the initial stresses s are


y0
0
2xy
xy
s = 0
x0 2

2xy 2xy
0

(2.109)

(2.110)

The stress stiffness matrix of a layered shell element is therefore


k =

n

j =1


 



GT sj G + GT sj G dV z j z j 1
V

(2.111)

Taking advantage of the definition of the forces per unit length in


equation (2.36), the initial stresses s can be replaced and the integration over
the volume converts to an integration over the element area.


(2.112)
GT N G + GT N G d A
k =
A

with


Nx
N =
Nx y

Nx y
Ny

(2.113)

and

Ny

N = 0

N
2xy

0
Nx

N
2xy

N xy
2
N
2xy

(2.114)

48

Fundamentals

Chapter 3

Sensitivities
The expression sensitivity origins from the so-called sensitivity analysis which
is a common methodology in the field of optimization with mathematical programming. The sensitivities express the influence of a change of the design
variables x to the objective function f . From the mathematical point of view,
the sensitivities are the gradients of an objective function. The general expression is therefore
f =

df
f (x + x) f (x)
= lim
dx x0
x

(3.1)

A simple method for the numerical determination of the sensitivities is the


finite difference approximation. The forward finite difference approximation is
defined as

f (x1 ,..., x i + x i ,..., x n ) f (x1 ,..., x i ,..., x n )


df
=
dx i
xi

(3.2)

For an objective function f with n design variables, n + 1 function evaluations


are needed to fully evaluate the sensitivities. In almost any problems of structural optimization, a finite element run is included in the evaluation of the
function. If the size of the finite element model is large or if many design variables are used, the numerical determination of the sensitivities may become
too expensive. Additionally, the numerical determination includes a numerical
error which is caused by the finite value of x i . However, a numerical determination of the sensitivities is the only possibility if the objective function is not
known explicitly, for example in black box optimizations.
Instead of evaluating the sensitivities numerically, they can be determined
with analytically derived formulas. In the first step, the objective function f

50

Sensitivities

is analytically differentiated with respect to the design variables. This is only


possible if the objective function is completely known and at least once differentiable. The evaluation of the resulting formulas, which is of course done with
numerical procedures, is straight forward and much faster than a finite difference evaluation. The presented method for the generation of reinforcement
doublers is based on a finite element model as well. The formulations of a finite
element framework are generally known wherefore an analytical derivation of
the sensitivities is possible. However, commercial finite element programs do
not provide their code (e.g. finite element formulations) explicitly. In order to
combine analytical sensitivities with finite elements, the finite element framework has to be implemented by itself.
The formulation of a general layered shell element has been presented in
Section 2.4.3. In the following sections, the analytical sensitivities for different objective functions are derived. The first section contains the sensitivity
derivations of stiffness, mass and stress stiffness matrices of a layered shell
finite element. Thereby, the thicknesses of the single laminate layers are
taken as design variables. Thus, the sensitivities provide an assumption of
the change of the objective function f for an infinitesimal change of a laminate
layer t within a finite element. Aiming for minimum weight structures, the objective function change d f with respect to the mass change dm are required.
Using the relation
dm = Adt

(3.3)

where denotes the density and A the shell element area, the sensitivities
with respect to the layer thickness can be transferred to the sensitivities with
respect to the layer mass.
1 df
df
=
dm A dt

(3.4)

Since the sensitivities are usually used for a relative guess, some simplifications can be made if all layers consist of the same material
1 df
df

dm
A dt

(3.5)

and if all elements additionally have the same area.


df
df

dm
dt

(3.6)

Based on these results for the system matrices, the sensitivities for eigenfrequency, minimal compliance, buckling and strength problems are derived.

3.1 Finite Element Based Sensitivities

51

Furthermore, a hybrid evaluation method is proposed which enables an optimization of arbitrary finite element models. Finally, a smoothing mechanism
is presented which reduces peaks in the sensitivity field.

3.1 Finite Element Based Sensitivities


3.1.1

Stiffness Matrix Sensitivities

The general formulation of the stiffness matrix of a layered shell element is


given by equations (2.99) to (2.105). The formulation is dependent on the positions of the laminate interfaces z j . To differentiate the stiffness matrix with
respect to the layer thicknesses t j , they have to be reformulated. The position
z j is related to the layer thicknesses with
zj =

j

k =1

t k z0

(3.7)

where z0 denotes the distance between the reference plane and the bottom of
the laminate (see Figure 3.1). Keep in mind that z0 is usually a function of

hn

zn

z0

zj
z1

hj
h1

h
2

reference-plane
laminate mid-plane

h
2

Figure 3.1: Schematic laminate layup with reference plane and laminate midplane
the layer thicknesses t k and thus not a constant. The reference plane coincides
with the mid-plane of the finite elements which is not necessarily equal to
the laminate mid-plane. Thus, a laminate offset can be controlled with z0 .
Reference plane and laminate mid-plane coincide if relation
z0 =

n
1 
tk
2 k =1

(3.8)

is fulfilled which implies no offset. It makes sense to postulate that for symmetric laminates. Using the same conventions as before where indices stand for

52

Sensitivities

membrane (m), coupling (c), bending (b) and transverse shear (s), the stiffness
matrix parts depending on the layer thicknesses t j are given by



!
j
j
1
n 


T
km =
Bm Q j Bm d A
t k z0
t k z0
(3.9)
A

j =1

kc =


n  

A

j =1

kb =

ks =

k =1

T
BT
m Q j Bb + Bb Q j Bm


n 


BT
b Q j Bb d A

j =1


n 

A

j =1

1
3

S
BT
s Q j B s d A





k =1

1
dA
2

k =1

k =1

j

k =1

t k z0

t k z0


j


j




t k z0

j
1
k =1

j
1
k =1

j
1
k =1
3

t k z0

(3.10)

t k z0

(3.11)
!

t k z0

(3.12)

Considering a bending loadcase, the change of the thickness of a middle layer


has not only an impact on the stiffness caused by itself. The changing layer
may also modify the stacking position of the surrounding layers which effects
another stiffness change. Using the convention above, the physical behavior of
a thickness change of a layer is captured completely.
In case of a linear integration as it is done for the membrane and the transverse shear part, the position of the reference plane z0 vanishes. This is comprehensible since the stacking position or the offset of the laminate, respectively neither have influence on the membrane nor on the transverse stiffness.
Taking advantage of relation
j

k =1

tk

j
1
k =1

tk = t j

(3.13)

the stiffness for the membrane and shear parts can be simplified to

n 

BT
Q
B
d
A
t
km =
j
m j m
j =1

ks =

n

j =1


A

S
BT
s Q j B s d A t j

(3.14)

(3.15)

Consequently, the sensitivities of these parts are only dependent on the material properties of the respective layer and yield

dkm
= BT
(3.16)
m Ql Bm d A
dt l
A

3.1 Finite Element Based Sensitivities

dks
=
dt l


A

53

BT
s Q l B s d A

(3.17)

For simplicity, it is assumed that the stiffness correction factor is independent


on the thickness t l .
The derivatives of bending and coupling parts are more complex because
the thickness appears with a higher order. In a first step, the bending stiffness
expression (3.11) is separated into three parts where the first part includes the
contributions of the subjacent layers of layer l, the second part the contribution
of layer l, and the third part the contributions of the overlying layers.
kb =

j =1


+
+



l
1

BT
b Q l Bb d A


n


j = l +1

1
3

BT
b Q j Bb d A

1
3





k =1

l

k =1

BT
b Q j Bb d A

j


t k z0

t k z0

1
3

j

k =1



l
1
k =1
3

t k z0

j
1
k =1

t k z0
3

t k z0


j
1
k =1

t k z0

(3.18)

This separation is important since the parts depend differently on the thickness t l of the considered layer l. The derivatives of the relation between thickness and interface position (3.7) are different whether a layer is contributing
or not.

z0

if j l

j


t l
(3.19)
t k z0 =

z0
t l k =1

if j < l

t l

The general expression for the bending stiffness sensitivities yields


kb
t l

l
1
j =1


+
+





BT
b Q j Bb d A


BT
b Ql Bb d A


n


j = l +1

l

k =1

j

k =1

2

t k z0
2

t k z0

BT
b Q j Bb d A



j

k =1

z0
t l

z0

t l
 

k =1

l
1

2

t k z0

2

z0

z0

!!

t l

t l

2

!!
j
1
z0
z0
1
t k z0
1

t l
t l
k =1

2

t k z0

  j 1

t k z0

k =1

(3.20)

54

Sensitivities

This equation can be simplified for special configurations of the laminate. Assuming that the reference plane coincides with the mid-plane of the laminate,
the offset is zero and z0 is defined as
z0 =

n
1 
tk
2 k =1

(3.21)

and the derivative is


dz0 1
=
dt l
2

(3.22)

The derivatives of the bending stiffness form to




2
2
  j 1
!!
j
1 


1
1
dkb l
T
=
Bb Q j Bb d A
t k z0
t k z0

dt l
2
2
A
j =1
k =1
k =1


2
2



l
l
1

1
1
t k z0

t k z0

+ BT
b Ql Bb d A
2
2
A
k =1
k =1



2
2 !!
j
j
1
n


1
1
T
Bb Q j Bb d A
t k z0

t k z0
(3.23)
+
2
2
A
j = l +1
k =1
k =1
Considering a symmetric laminate, the derivatives can be simplified significantly. Using the parameterization above, a symmetric laminate can simply be
represented by setting z0 to zero, which implies that the reference plane coincides with the bottom of the laminate, and simultaneously double the stiffness
contributions to take into account that only half of the laminate is considered.

2

l

dkb
T
= 2 Bb Q l Bb d A
tk
dt l
A
k =1


2  j 1
2 !!
j
n



T
+2
Bb Q j Bb d A
tk
tk
(3.24)
j = l +1

k =1

k =1

Here it becomes obvious that a change of the bending stiffness is caused by two
different effects. The first term expresses the stiffness change due to the thickness change of the layer itself. The terms within the summation consider the
change of the stiffness caused by pushing outward the overlaying layers which
results in a higher area moment of inertia (see Figure 3.2). This relation can be
written alternatively with expressions (3.25) and (3.26). The top layer (l = n) is
independent on the subjacent layers wherefore its sensitivity is expressed with

2

n

dkb
T
= 2 Bb Q l Bb d A
tk
(3.25)
dt l
A
k =1

3.1 Finite Element Based Sensitivities

55

z3
z2 k3(z2+t,z3+t)
z1 k2(z1+t,z2+t)

k3(z2,z3) 3
z
k2(z1,z2) 2
z
k1(z0,z1) z1

z0

k1(z0+t,z1)

Figure 3.2: Effect of a thickness change to the overlaying layers

The lower layers (l = 1,..., n 1) take into account the stiffness contribution of
the overlaying layers wherefore

dkb
=2
dt l


A

BT
b Ql Bb d A


A

BT
b Q l +1 B b d A



l

k =1

tk

dkb
dt j +1

(3.26)

This convention has been used in [145]. It may be advantageous for the implementation since a summation can be saved.
The sensitivities for the coupling parts are derived analogously to the bending parts. An explicit derivation is given in Appendix B. The material stiffness parts and the strain-displacement matrix are replaced corresponding to
equation (2.100) and the exponent is reduced by one order. Keep in mind that
the coupling stiffness part vanishes for a symmetric laminate.
The total sensitivities of the stiffness matrix are built with an addressed
summation of the several parts corresponding to equation (2.105).

dk dkm dk c dkb dks


=
+
+
+
dt l
dt l
dt l
dt l
dt l

(3.27)

All the derivations above have been performed on the element level for the
element stiffness matrix k. The sensitivities of the global stiffness matrix K
can however be derived directly from the element stiffness parts. The global
stiffness matrix is an addressed summation of all the element matrices which

56

Sensitivities

is denoted schematically in equation (3.28).

K=

k1
k2
k3

..
.

(3.28)

The element stiffness matrix derivatives are dependent only on the thickness of
the respective element. The derivatives with respect to the thicknesses of other
elements are all zero. Thus, the global stiffness matrix derivative dK
d t l contains
only zeros except of the entries corresponding to the considered element l which
is shown schematically in equation (3.29).

dK
=
dt l

dkl
tl

(3.29)

3.1.2 Mass Matrix Sensitivities


The lumped mass matrix of a layered shell element is given in equation (2.106).
The layer thickness t j occurs only linear wherefore the derivative is simply

dm
= l T d A
(3.30)
dt l
A

Analogously to stiffness, the element mass matrices m are only sensitive to


a thickness change of themselves. The derivatives of the global stiffness
matrix M are assembled according to the scheme shown in equation (3.29).

3.1 Finite Element Based Sensitivities

3.1.3

57

Stress Stiffness Matrix Sensitivities

As shown for stiffness and mass matrix, a change of a layer thickness only affects the sensitivity of the respective element. This is however not the case for
the stress stiffness matrix. According to the definition given in equation (2.87),
the stress stiffness matrix is dependent on the stresses within the element.
A change of a layer thickness of one element may change the stress field of
the entire structure. Thus it has not only an influence on the stress stiffness
matrix of the considered element but on the entire model.
The stress stiffness matrix for a layered shell element is given in
equation (2.112). Since the shape function derivatives G are not dependent on
the element thickness, the derivatives of the geometric stiffness matrix with
respect to the thickness of the j th layer of the i th element is
dk
=
dt i j



dN
dN
GT
G + GT
G dA
dt i j
dt i j
A

(3.31)

For the derivation of the sensitivities, the line loads Nk of the k th element are
considered. Notice that the index k does denote the element number and not
a component of the load line. It is defined as the sum of the products of the
respective stress component and layer thicknesses for all layers l.
Nk =

l

j =1

k j t k j

(3.32)

Thus, the line load of the k th element is a function of all stresses within the
element and the corresponding thicknesses.
Nk = Nk ([t1 ,..., t l ] ,[1 (1 (u (t))),..., l ( l (u (t)))])

(3.33)

Nk is not only dependent explicitly on the layer thicknesses of the element, but
also on the layer stresses which are a function of the displacement field and
therefore on the layer thicknesses again. Using the chain rule, the line force
derivatives of the k th -element with respect to the j th layer of the i th element
yield
 
!
lk

Nk k j k j
u
dN k Nk
=
+
(3.34)
dt i j
t i j
t i j
j =1 k j k j u
Through the displacement field, a change of an element layer thickness has
an impact on the stress stiffness matrices of any element even if these contributions might be small. The partial derivative of the line load component

58

Sensitivities

with respect to a layer thickness is nothing else than the layer stresses k j assuming that the considered thickness t k j is member of the respective element.
Nk
t i j

"
=

k j
0

if
if

i=k
i = k

(3.35)

The differentiation of the line load component with respect to the layer stresses
yields the layer thicknesses t k j . The derivative of the stresses with respect
to the strains is per definition the material stiffness matrix Ck j . The strains
are connected to the displacements with the strain-displacement matrix Bk j .
Thus, equation (3.34) is simplified to

!
lk 

 u
dN k Nk
=
+
t k j Ck j Bk j
(3.36)
dt i j
t i j
t i j
j =1
The components of this result can be employed for the stress stiffness matrix
sensitivities in equation (3.31).

3.1.4 Displacement Sensitivities


The derivation of the displacement sensitivities starts from the fully assembled structure of the static finite element problem which has been derived in
Section 2.4.1.
Ku = r

(3.37)

As before, K denotes the stiffness matrix, u the displacement vector and r the
corresponding load vector. By rearranging the rows and columns of the linear
system, the formulation can be partitioned into

   
K11 K12 u x
rc
(3.38)
=
K21 K22 u c
rx
where subscript c denotes known quantities and subscript x denotes unknown
quantities. On every degree of freedom either the displacement u or the external force r but not both have to be defined. Where loads r c are prescribed, the
corresponding displacements u x are unknown and vice versa. Keep in mind
that no external force, or r = 0 respectively, is also a prescribed boundary condition. Typically, r x represents the support reaction forces. The first row of the
system represents the reduced equation system where the boundary conditions
are already introduced.
K11 u x = r c K12 u c

(3.39)

3.1 Finite Element Based Sensitivities

59

In contrast to the general stiffness matrix K, the reduced stiffness matrix K11
is not singular if u c determines an at least statically determinate support and
the linear system becomes solvable.
The system is simplified again if only homogeneous geometric boundary
conditions (u c = 0) are used.
K11 u x = r c

(3.40)

The total derivatives with respect to the design variables t are given by

d K11
dt K21

K12
K22

  
ux
K11
+
uc
K21


 
 
d rc
K12 d u x
=
K22 dt u c
dt r x

(3.41)

We assume that prescribed displacements u c are independent of the design


variables.
du c
=0
dt

(3.42)

Equation (3.41) can thus be simplified to



d K11
dt K21

K12
K22

  
ux
K11
+
uc
K21

K12
K22

  dux 
dt


=

drc
dt
drx
dt

The derivatives for the unknown displacements are




dK11
dK12
du x
1 dr c
ux
uc
= K11

dt
dt
dt
dt

(3.43)

(3.44)

The general displacement derivative expression is composed with


equations (3.42) and (3.44).
If only homogeneous geometric boundary
conditions are applied, which means that the prescribed displacements are all
zero (u c = 0), equation (3.44) simplifies to


dK11
du x
1 dr c
ux
= K11

(3.45)
dt
dt
dt
The prescribed loads may be dependent on the thickness variables, for example
if gravitation is taken into account. The general load vector for body forces
is defined at the right side of basic finite element equation (2.69). In case of
gravitation it yields

(3.46)
r = g dV
V

60

Sensitivities

where denotes the density and g the gravitational constant. The derivatives
are thus simply

dr
(3.47)
= g d A
dt
A

For many structural design problems, the prescribed loads are design independent whereby equation (3.45) simplifies to
du x
1 dK11
ux
= K11
dt
dt

(3.48)

The evaluation of the displacement sensitivities is equivalent to solving a linear system equation with multiple right hand sides. The right hand side is
given by the term in parenthesis in equation (3.44). It is built by the multiplication of the stiffness matrix sensitivities and the displacement vector u. The
evaluation of the right hand term is thus fast. However, it has a dimension
of number of degrees of freedom wherefore the solution of the linear equation
system is relatively expensive. The resulting displacement derivatives form a
matrix with dimension of number of degrees of freedom times the number of
design variables which is fully occupied for the most cases. With increasing
model size and/or number of design variables, the required memory becomes
huge or even critical. Displacement sensitivities should be used only when
needed, e.g. buckling load factor or strength sensitivities.

3.2 Eigenfrequency Sensitivities

61

3.2 Eigenfrequency Sensitivities


The finite element equation for harmonic vibration problems has been derived
in Section 2.4.1 and is given by equation (2.74). Considering only the n th
eigenvalue n and the corresponding eigenvector n , the expression forms to
(K n M) n = 0

(3.49)

Thereby, K denotes the global stiffness matrix and M the global mass matrix. The eigenvalue n is the square of the angular frequency n (corresponding to equation (2.74)) which is 2 times the frequency f n . Thus, performing
a maximization of the eigenvalues n is equivalent to a maximization of the
frequency f n . The total derivative of the harmonic problem can be expressed
as
d
((K n M)n ) = 0
dt

(3.50)

and the expanded form is




dK d k
dM
d n
M n

n + (K n M)
=0
dt
dt
dt
dt

(3.51)

This expression can be simplified by multiplying the individual terms with the
transpose of the eigenvector T
n from the left side.


dK d n
dM
d n
M n
Tn

n + Tn (K n M)
=0
(3.52)


 dt
dt
dt
dt
=0

The second term of the expression vanishes applying the initial equation (3.49).
A rearrangement of the terms leads to the general equation for the eigenvalue
sensitivities which is


dM
T dK
d n n dt n dt n
=
(3.53)
dt
Tn Mn
This expression is generally accepted and has been derived and applied in
several publications (e.g. [185, 26, 108, 103, 128]). The entries of the
eigenvectors n are relative so that they can be M-orthonormalized. Consequently, the generalized mass which is defined as T
n Mn becomes equal to 1
wherefore the denominator of equation (3.53) can be removed. The simplified
sensitivity equation for the orthonormalized eigenvectors forms to


dK
dM
d n
= T

n
(3.54)
n
n
dt
dt
dt

62

Sensitivities

Seyraniant et al. [152] demonstrated that this expression is valid for single
eigenvalues only. Multiple eigenvalues may for example occur for quadratic
plates under ideal conditions. However, such cases are implausible for real applications wherefore it is assumed here that all eigenvalues are single. The
sensitivities above have been derived with respect to the element layer thickness t. However, expression (3.53) is universally valid and the thickness array
could for example be replaced with the array of the orientation angles .
To reduce the calculation costs, the evaluation is performed on an element
level. As shown in Section 3.1.1, entries excepting the ones related to the considered element are zero. Having rows and columns with only zero entries, a
multiplication with the eigenvector n becomes unnecessary. Instead, the element matrix derivatives are multiplied directly with the respective entries of
the eigenvector e n . The number of calculations steps is reduced drastically
and the evaluation is much more efficient.


dm e
d n e T dk
= n
n
n
dt
dt
dt

(3.55)

Both, element stiffness matrix k and element mass matrix m, the global
matrices K and M as well as their derivatives are positive-definite. Thus, the
result of the both sided multiplication with the eigenvector n must be positive. Due to the minus sign within the braces of equation (3.54), the sensitivities may however become negative. Negative gradients indicate areas where
material must be reduced to increase the objective function which is in this
case the eigenvalue or the frequency respectively. Aiming for mass minimal
structures that fulfill the stiffness requirement, the reduction of material in
these areas is essential. The occurrence of negative gradients can also be explained physically. The existence of material affects always the stiffness and
the mass. Generally, the stiffness increases the frequencies of the structure
and the mass causes the frequencies to drop. More decisive than the existence
of stiffness and mass is its distribution. Mass located in regions with high
amplitudes, which occurs usually far away from bearings, causes a higher frequency drop. On the other hand, mass close to bearings does not affect the
frequencies significantly. Considering harmonic problems, the engineer must
place the material in a way that the stiffness contribution is higher than the
mass penalty in order to achieve maximal frequencies with minimal mass. If
the mass penalty is higher than the stiffness contribution, which is expressed
with
n T
n

dM
dK
n > Tn
n ,
dt
dt

(3.56)

3.2 Eigenfrequency Sensitivities

63

the sensitivities become negative because the existence of material is counterproductive. Identifying these areas intuitively is almost impossible. Thus,
using the sensitivity field may help to find improved designs.
The sensitivities of a harmonic vibration problem are illustrated on the example of an all-side clamped CFRP-plate. The plate has an aspect ratio of 2
and is made of an unidirectional ply-material (see Appendix Table A.2) with
orientation angles (0/45/-45/90)S where the 0 -direction is parallel to the long
side. Figure 3.3 shows the first eigenmode of this plate. The sensitivities

y
x
Figure 3.3: First eigenmode of a all-side clamped plate
for the symmetric half of the composite are shown in Figure 3.4 (a/c/e/g). Consider that the color scales of the four pictures are different. Red areas indicate high sensitivity values whereas the thickness of the respective layer in
these areas should be increased in order to raise the first eigenfrequency maximally. The maximal sensitivities occur on the long side edge of the 90 -layer
which seems to be plausible. In contrast, the blue areas indicate low sensitivity. The threshold between positive and negative sensitivities is emphasized
in Figure 3.4 (b/d/f/h). Blue areas have negative sensitivities wherefore material should be removed in order to increase the frequency. Obviously, a big
area of the 0 -layer material is useless or even counterproductive. However,
also the other layers hold locations where material could be removed without having a frequency loss. The positive-negative plots in Figure 3.4 may
give the wrong impression that material can be removed completely in bluecolored areas. However, the sensitivities are calculated for the mode shown
in Figure 3.3. Removing the material in one step would lead to a completely
different eigenmode wherefore the wrongly assumed optimal material distribution becomes sub-optimal. The sensitivity field is calculated for one specific
mode and changes when the design is modified. In order to find an optimal
thickness distribution, the model should be adjusted step-wise. After each
small modification, the sensitivities have to be reevaluated considering the new
eigenmode.

64

Sensitivities

(a) = 0

(b) = 0 ()

(c) = 45

(d) = 45 ()

(e) = 45

(f) = 45 ()

(g) = 90

(h) = 90 ()

Figure 3.4: Regular sensitivities (a,c,e,g) (not same scale) and -plot of the
sensitivities (b,d,f,h)

3.3 Minimal Compliance Sensitivities

65

3.3 Minimal Compliance Sensitivities


The compliance of a structure or the stiffness, respectively, is usually expressed
with help of the potential of the external forces W. It is built by the multiplication of the displacement vector u and the load vector r. Alternatively, the load
vector can be replaced with the stiffness matrix K applying equation (3.37).
Analogously to the derivation of the displacement sensitivities, the degrees of
freedom are partitioned in to the prescribed ones, indicated with c, and the free
ones indicated with x, respectively.
 T  
ux
rc
uc
rx
 T 
ux
K11
= uT Ku =
uc
K21

W = uT r =

(3.57)
K12
K22

 
ux
uc

(3.58)

Due to the positive-definiteness of the stiffness matrix, the theoretical minimum of the potential of the external forces W is zero. It is reached for an
infinitely stiff structure where all displacements are zero. However, this is
only valid for a completely force-controlled loadcase where the prescribed displacements are all zero (u c = 0) and the deformation field results of the applied
external forces. Considering a displacement-controlled loadcase, the displacements become never zero and the minimal potential of the external forces is
only achieved if the corresponding reaction forces become zero. This is the case
for an infinitely soft structure wherefore the potential of external forces is not
feasible anymore to express the stiffness. Moreover, from the engineering point
of view, it is not clear what "stiff" means if displacements are prescribed.
Consequently, the following derivations and calculations for minimal compliance structures consider only completely force-controlled loadcases. The
sensitivities of the potential of external forces W are derived applying the chain
rule to equation (3.58).

dK
du
dW
=u T
u + 2uT K
dt
dt
dt
 T

 
d K11 K12 u x
ux
=
uc
dt K21 K22 u c
 T 

 
ux
K11 K12 d u x
+2
uc
K21 K22 dt u c

(3.59)

(3.60)

66

Sensitivities

The prescribed displacements u c are fixed so that their derivatives are zero.
 T

d K11
dW
ux
=
uc
dt
dt K21
 T
ux
K11
+2
uc

K12
K22
K12

 
ux
uc

du x

dt

(3.61)

With the assumption of a completely force-controlled loadcase (u c = 0),


equation (3.61) simplifies to
dK11
du x
dW
u x + 2uTx K11
= uT
x
dt
dt
dt

(3.62)

By substituting the unknown displacement derivatives with the above derived


sensitivity formula (3.45), further simplification can be made.


dW
dK11
T dK11
T
1 dr c
u x + 2u x K11 K11
ux
= ux

(3.63)
dt
dt
dt
dt
dK11
dr c
u x + 2uTx
(3.64)
= u T
x
dt
dt
Since the prescribed displacements u c are assumed to be zero, they do not
contribute and the expression can be generalized to
dK
dr
dW
u + 2uT
= u T
dt
dt
dt

(3.65)

Referring to the definition of the potential of external forces in equation (3.58),


a structure holds maximal stiffness if W is minimal. If high sensitivity values should indicate areas which have to be reinforced in order to increase the
stiffness, equation (3.65) must be negated. For the following explanations and
also to be consistent with the eigenfrequency sensitivities in Section 3.2, the
negative potential W which is indicated with an overline is introduced.
dK
dr
dW
u 2uT
= uT
dt
dt
dt

(3.66)

Using this general expression, sensitivities may become negative. Considering


a loadcase with gravitation there may exist regions where additional material
increases the weight more than it contributes to the stiffness. Assuming that
loads are design independent, the above expression simplifies again to
dK
dW
= uT
u
dt
dt

(3.67)

3.3 Minimal Compliance Sensitivities

67

Due to the positive definiteness of the stiffness derivatives, load independent


minimal compliance sensitivities are strictly positive. It is not possible to reduce the stiffness by adding material to any location of the structure. Comparing equation (3.67) and the eigenfrequency sensitivity equation (3.54), some
similarities can be noticed. Both sensitivities are built with the quadratic product of deformation vector and stiffness matrix derivatives whereas the eigenvalue sensitivities additionally contain the mass matrix sensitivities. However,
assuming that the eigenvalue is going towards zero, the eigenfrequency sensitivities yield

dK
d n
= T
n
n
dt
dt

(3.68)

which is identical to equation (3.67) disregarding that the deformation is defined differently. The minimal compliance sensitivities are dependent on the
loading of the structure which is given by the right hand side of equation (2.71).
Thus, these sensitivities can be used to improve the stiffness of a structure with
respect to a specific loadcase. In contrast, no load is defined for the eigenfrequency calculation and they are simply dependent on the displacement boundary conditions.
The two different types of minimal compliance sensitivities are demonstrated on a structure which is inspired by a bridge. The design space is defined with a rectangular with side ratio of 4:1 as shown in Figure 3.5. The

y
x

Figure 3.5: Design space and loads of the bridge structure


structure is simply supported on the left and the right ends and a uniform load
is applied on the top. Again we consider a layup of unidirectional reinforced
plies with orientation angles (0/45/-45/90). In the first case, the sensitivities
are calculated assuming constant loads using equation (3.67). The according
sensitivity fields are plotted on the right hand side of Figure 3.6 where the
colorbar is valid for all four plots. Consider that all sensitivities in that case
are strictly positive. In the second case shown on the left side of Figure 3.6, a
gravitational field is added. Thus, the loads become dependent on the design
variables wherefore the sensitivities are evaluated with equation (3.66). The
two sensitivity fields seems to be almost identical. However, the colorbar for

68

Sensitivities

the case with gravity indicates that some elements have negative sensitivities
which implies that a thickness increase would increase the displacements of
nodes at which external forces are applied due to the additional weight forces.

(a) = 0

(b) = 0 with 
g

(c) = 45

(d) = 45 with 
g

(e) = 45

(f) = 45 with 
g

(g) = 90

(h) = 90 with 
g
x10

10

15

-3

20

-0.02

0.02

0.04

0.06

0.08

Figure 3.6: Sensitivities of the bridge structure with and without gravity

3.4 Sensitivities of Nodal Displacement


Instead of optimizing the stiffness with the potential of external forces, the
nodal displacements can be considered separately. A general objective function
to do so may have the following form.

T 

x+u = T (x + u) w T (x + u)

(3.69)

The vector x contains the nodal positions and u the displacements. The matrix is used to select the degrees of freedom involved in the optimization.

3.4 Sensitivities of Nodal Displacement

69

The diagonal matrix w contains weighting factors. Consider a four-noded


membrane element under uniaxial tension in x-direction shown in Figure 3.7.
The degrees of freedom in the two in-plane directions are denoted with u =

v4

v1

v3

u4

v2

u1

u3

u2

Figure 3.7: Four-noded element


T
u1 v1 u2 v2 u3 v3 u4 v4 . A potential objective could be the minimization of the distance between nodes 1 and 2. Thus the vector is
=


1
0

0
1

1
0

0
1

0
0

0
0

0
0

T
0
0

(3.70)

and the weighting matrix is


w=


1
0

0
1

(3.71)

The objective function then yields




= ((x1 + u1 ) (x2 + u2 ))2 + ((y1 + v1 ) (y2 + v2 ))2

(3.72)

The function reaches its maximum if nodes 1 and 2 coincide for the deformed
case. Alternatively, the relative displacement between two nodes can be optimized taking advantage of the proposed objective function. The nodal position
vector x is set to zero and only the displacements are considered.

T 

u = T u w T u

(3.73)

Using the same and w as above, the objective function is




= (u1 u2 )2 + (v1 v2 )2

(3.74)

The minimal value is reached if the relative displacement between nodes 1


and 2 is zero which implies that no strain occurs between these two nodes.

70

Sensitivities

Such a minimization could be helpful when aiming for structures with good
dimensional stability.
The derivative of the general objective function is

T  du 
d
= 2 T (x + u) w T
(3.75)
dt
dt
The main drawback of these sensitivities is that displacement derivatives are
needed. As mentioned in Section 3.1.4, their calculation is expensive and
should be avoided if possible.
The sensitivity field of nodal displacements is illustrated on an example
of a grabber. The design space is shown in Figure 3.8(a). It is symmetrically
loaded on the left side with a unit force. For an isotropic material, the reference
points P re f on the left side drift apart as shown in Figure 3.8(b). A potential
optimization goal could be to reduce the drifting of the two reference points or
even to make them approaching each other. This can be formulated with an
objective function as introduced above. A specific utilization of unidirectional
laminate plies can change the deformation field significantly. Such effects are
used for compliant mechanisms which are able to do large deformations without hinges. This example is further developed in the application Chapter 6.2.
It is clear that the structure cannot be designed as specific when using the min-

Pref

y
x

Pref

(a) Design space

(b) Displacement field

Figure 3.8: Grabber example with a homogeneous, isotropic material


imal compliance, namely the potential of external forces as objective functions
since it is global and considers the displacement field of the entire structure.
The sensitivities when using a laminate of unidirectional plies with orientation
angles of (0/45/-45/90) is shown in Figure 3.9.

3.5 Buckling Load Factor Sensitivities

71

(a) = 0

(b) = 45

(c) = 45

(d) = 90

Figure 3.9: Nodal displacement sensitivities of the grabber example

3.5 Buckling Load Factor Sensitivities


The basic finite element equation for the bifurcation buckling problem, which
has been derived in Section 2.4.2, is similar to the harmonic problem (3.49) and
also solved with an eigenvalue analysis.
(K + n K ) n = 0

(3.76)

Instead of the mass matrix M, the stress stiffness matrix K is used. The
eigenvalues n describe the buckling mode whereas the eigenvalue n is the
respective buckling load factor which quantifies the potential load increase until the buckling occurs. Buckling load sensitivities can be found by applying

72

Sensitivities

the same operations as done in equations (3.50-3.53).




dK
T dK
d n n dt + n dt n
=
dt
Tn (K ) n

(3.77)

Since the eigenvalue problem for buckling contains a plus sign instead of a
minus sign, the negative of the stress stiffness matrix stays in the denominator.
Analogously to the eigenfrequency sensitivities, the expression T
n (K ) n is
sometimes called the generalized mass, even if it is here not related to mass.
Performing a K -orthonormalization simplifies the expression to


dK
dK
d n
= T
+

n
(3.78)
n
n
dt
dt
dt
The current sensitivity equation for the buckling load factor is valid for the
global system while the sensitivities of the stress stiffness matrix
dk,k

dt


=

GT

dN k
G dV
dt

(3.79)

are derived on element-level. An evaluation of the buckling load sensitivities


is computationally inefficient if done according to equation (3.78) which is in
agreement with the investigations of Mateus et al. [109]. The computational
costs can be reduced by evaluating the sensitivities on an element-level. Consequently, equation (3.78) is replaced with an addressed summation of the subparts which yields




elm
dk,i
dk i
d n n
=
Tn,i
+ n
n,i
dt
dt
dt
i =1

(3.80)

Nevertheless, the numerical costs for evaluating the stress stiffness sensitivities with the equations above are still high. Especially for large models, the
dependency of the elements on the displacement field causes many additional
calculation steps.

3.5 Buckling Load Factor Sensitivities

3.5.1

73

Simplifications

In order to reduce the computational costs for the evaluation of the buckling
load factor sensitivities, it might be helpful to make some simplifications.

P
P

P
(a) serial loading

(b) parallel loading

Figure 3.10: Element cluster with serial and parallel loading


Assuming a finite element model where the elements are perfectly serially
loaded according to Figure 3.10(a), the load P is not dependent on the element
thickness and consequently constant.
P = const.

(3.81)

The stresses in the j th layer are given by


j =

P
tjbj

(3.82)

Recalling the definition of the line loads


N=

l

j =1

j tj

(3.83)

and taking their derivatives, it becomes obvious that they are independent on
a change of the element thicknesses.
N
t j

j
t j

tj +j =

tj +
bt2j

P
=0
bt j

(3.84)

Consequently, the derivatives with respect to the thickness become zero wherefore the stress stiffness matrix derivatives yield zero as well. The sensitivities
of the buckling load factors which are given in equation (3.78) are then simplified to

dK
d n
= T
n
n
dt
dt

(3.85)

74

Sensitivities

which reduces the computational costs drastically since the stress stiffness matrix has not to be considered anymore.
Taking a connection where the elements are loaded in parallel as shown in
Figure 3.10(b), the load P is now dependent on the thickness of the particular elements wherefore the stress stiffness matrix sensitivities are not zero
anymore and cannot be neglected. For most cases, the finite element models
contain a combination of serial and parallel connected elements wherefore the
simplification (3.85) cannot be employed without accepting an error. However,
it might be feasible to neglect the stress stiffness term if its contribution is
much smaller than the contribution arising from the stiffness matrix which
can be expressed with

Tn

dK
dK
n  n Tn
n
dt
dt

(3.86)

If the contribution of the stress stiffness matrix is very small, it is unreasonable to calculated its expensive sensitivities, especially if used in a preliminary
design process, and the equation for the buckling load factor sensitivities is
simplified to

dK
d n
= T
n
n
dt
dt

(3.87)

Consider an isotropic plate in Figure 3.11(a) which is fully clamped on the left
side and simply supported on the other side. The plate is loaded with an uniform pressure load on the right hand side which leads to the first buckling
mode shown in Figure 3.11(b). For such a case, it can be assumed approximately that the elements are connected serially. Figure 3.12(a) represents the

(a) Plate configuration with load

(b) First buckling mode

Figure 3.11: Example of a buckling plate


sensitivity field for a full evaluation using equations (3.80) and (3.36). Sensitivities are high in areas where the virtual bending moments are high. The

3.5 Buckling Load Factor Sensitivities

75

sensitivity field where only the stiffness matrix derivatives are taken into account, according to equation (3.87), is shown in Figure 3.12(b).
400
350
300
250
200
150
100
50

(a) Full evaluation


450
400
350
300
250
200
150
100
50

(b) Evaluation with stiffness matrix only

Figure 3.12: Buckling load factor sensitivity fields


It becomes obvious that the simplification is feasible for the current example. The deviations of the simplified evaluations are small. Keeping in mind
that the sensitivities are used for a rough design process and that the modeling errors may be much higher, using the simplified evaluation techniques is
justifiable. However, the feasibility of these simplifications has to be proved for
each example separately.

76

Sensitivities

3.6 Strength Sensitivities


In order to employ the proposed optimization procedure, the gradient field that
expresses the influence of an element-layer thickness change on the objective
function is needed. This presumes that the objective function is at least once
differentiable. In contrast to the potential energy, eigenfrequencies or buckling
load factors, there is no closed-form expression that characterizes the strength
of a structure since this is a very local property. The strength is limited by the
occurring stresses or strains at the critical region. Applying a design change
may cause a sudden relocation of it, rendering the problem character discontinuous. Considering a laminated composite structure, the structural strength (or
alternatively the critical region) is given by the maximal failure index. Using
a max-function results in a non-differentiable objective function.
In order to avoid using max-functions, a Pseudo Strength Function (PSF)
formulated in equation (3.88) will be used. It is inspired by the optimality
criterion called fully stressed design [113, 142, 52] which postulates that the
material distribution is optimal in terms of minimal weight when the material stresses are critical at every point of the structure. Having such a stress
state, no more material can be removed without causing failure. Requiring a
fail-safe structure, the safety factors, or in case of laminated composites, the
failure indices, must have a constant margin against failure for every material element. Such a stress state can usually be reached for truss structures
or beams. Considering a laminated structure, fully stressed designs can only
be found for simple geometries and homogeneous loadcases. However, this optimality criterion can be applied to find solutions that come closer to a fully
stressed design. Kress employs a similar approach optimizing the shape of a
flywheel [87] and sandwich onserts [88].
f TH =

mi #
$2
n 1 
1
F I i, j F I
n i =1 m i j =1

(3.88)

F I i, j indicates the failure index of the j th layer of the i th element and F I is the
average of all failure indices in the considered design space which can also be
expressed with equation (3.89).

mk
n
1 
1 
FI =
F I k,l
(3.89)
n k =1 m k l =1
The PSF can be understood as the variance of the failure indices of all m layers of all n elements. This function is differentiable when the applied failure
criterion is differentiable. The minus sign in front converts the minimization

3.6 Strength Sensitivities

77

problem to a maximization problem. Thus, areas where the laminate should


be reinforced in order to increase the strength have high gradient values. The
PSF is maximal (which is equal to zero) if the failure index variance of the
structure is zero. It is equivalent to a state where the material failure probabilities are distributed homogeneously over the whole design space which can
be interpreted as a fully stressed design. It can be expected that the optimal value of the function may never be reached in a real structural optimization problem. Moreover, it is not guaranteed that all failure indices fall below
the critical value of one. However, the formulation transforms the structural
strength which is limited by local failure into a global expression that can be
used as an objective function for optimization.

3.6.1

Pseudo Strength Function Sensitivities

The PSF sensitivities with respect to the element layer thicknesses can be built
by using the chain rule. They are given by


!
mi
 dF I i, j dF I

n 1 
1
d f PSF
=
2 F I i, j F I

(3.90)
dt
n i =1 m i j =1
dt
dt
where the derivative of the average failure index F I is defined by

m k dF I
n
1 
dF I 1 
k,l
=
dt
n k=1 m k l =1 dt

3.6.2

(3.91)

Failure Criteria

In order to ensure the differentiability of the pseudo strength function, an


appropriate failure criterion for unidirectionally reinforced fiber composites
has to be chosen. The stress based global failure criteria of Tsai-Hill and
Tsai-Wu are formulated with a function () which is quadratic in terms of
the stresses (see Section 2.3). They are capable of indicating a damage event
but they do not give any information about the failure mode. Their closed form
expressions make them predestined for an application in the proposed optimization method. Both criteria can be rated as rather conservative and may
be inappropriate when the design of light-weight structures should be pushed
to the limit. However, the applied design process which will be introduced in
Section 4 has been made for a preliminary design. The discretization of the reinforcements doublers due to the element-based formulation may have a larger
impact on the uncertainties than the used failure criterion. Thus, it is assumed
that these failure criteria comply with the requirements of the design process.

78

Sensitivities

In the following sections, the criteria of Tsai-Hill and Tsai-Wu are converted
to a strain-based matrix formulation. This simplifies the derivation of the failure index derivatives which are required for the sensitivities of the PSF in
equation (3.90) and makes their evaluation much more efficient.
Matrix Formulation of Tsai-Hill Criterion
The failure condition of the Tsai-Hill criterion [57, 166] is given by
 2
 2  2
1
1 2
2
12
( ) =

+
+
=1
2
X
Y
S
X

(3.92)

1 , 2 and 12 are the plane-stress components in material principal coordinates while X , Y and S are the corresponding material strength values. X and
Y has to be chosen depending on the stress state. If 1 is positive, X is set to
the tension stress limit in fiber direction X t and if it is negative, the compression stress limit X c is taken. The choice of Y is done analogically. The criterion
can be transformed into a matrix formulation
T
() = 12
STH 12

with the stress vector



1
12 = 2
12
and a stress-based strength matrix
1

2X1 2
0
X2
1

1
0 .
S
TH = 2X 2
Y2
1
0
0
S2

(3.93)

(3.94)

(3.95)

This notation simplifies the derivation of the sensitivities and its implementation. Since the strength matrix S
consists only of material strength data,
TH
it is not dependent on the element layer thicknesses. Only the stress vector is
sensitive to a design change what makes the differentiation simple. In contrast
to the element stresses in material principal coordinates 12 , which have to be
calculated in the post-processing, the element strains in reference coordinates
x y are a direct result of the finite element analysis. Therefore, equation (3.93)
is converted into a strain dependent expression by applying a couple of transformations. For that purpose, the material principal strains 12 are mapped to
the corresponding stresses 12 applying the material stiffness matrix Q.
12 = Q 12

(3.96)

3.6 Strength Sensitivities

79

In a second step, the global material strains are transformed to the material
principal strains by applying a rotation matrix T (2.26) and a so-called Reutermatrix R (3.97).
12 = RT R 1 x y

(3.97)

Combination of equations (3.96) and (3.97) leads to the expression (3.98).


12 = QRT R 1 x y

(3.98)

This expression is usually used to calculate the stresses in the layers based on
the element strains arising from the finite element analysis. However, instead
of applying the transformation to the element strains, it can be used to transform the stress-based strength matrix into a strain based formulation. It is
carried out by a both-sided multiplication of the stress-based formulation S
TH
with the derived transformation (3.98).

T
1
STH = QRT R 1 S
TH QRT R

(3.99)

The strain-based formulation enables an evaluation of the failure indices using


the global element strains x y .

() = Txy STH x y .

(3.100)

Using the stress based formulation, the transformation (3.98) has to be carried out for every layer of every element. This is an unnecessarily high effort
especially when done in an iterative optimization process. The stress-based
, the transformation components Q, R and T and thus,
strength matrix S
TH
also the strain-based strength matrix STH are unique for a layer covering any
elements. Using the strain-based formulation, the strength matrix has to be
evaluated only once for all layers. It does not change during the optimization.
The failure indices can be determined employing the element strains. They
are a direct result of the nodal displacement u of the finite element evaluation
multiplied with the matrix B i containing the element shape functions i of
the i th element. Considering only in-plane deformations, the element strains
are unique for all layers of an element. For loadcases with bending, the strains
have to be calculated with equation (2.89) which leads to


x y,i j = B m,i + z j B b,i u
(3.101)
The choice of the position z j where the strains are evaluated has to be take by
the designer. Due to the linear distribution of the strains, the maximal values occur at the interface which with maximal distance to the neutral axis.

80

Sensitivities

However, also the mid-plane of each layer can be taken which provides an average value of the strains. The failure indices F I can be expressed combining
equation (2.51) and (3.100). For clarification, the equation is indexed with element and layer numbers i and j, respectively.


(3.102)
F I i, j = + i, j = + Txy,i j STH j x y,i j .

Using expression (3.102), the derivatives of the failure indices are given by
dF I i, j
dt

d xy,i j

T
x y,i j STH j dt
= 
+
T
S
x y,i j TH j x y,i j

(3.103)

The evaluation of equations (3.90) to (3.103) is straight forward. However, obtaining the element strain derivatives is more expensive. They are depending on the nodal displacement derivatives as it is shown by differentiating
equation (3.101).
d x y,i j

dt


 du
= B m,i + z j B b,i
dt

(3.104)

The nodal displacement derivatives have been derived in Section 3.1.4.


Matrix Formulation of Tsai-Wu Criterion
The similar Tsai-Wu criterion (3.105) [169, 94] can be used analogically. In
contrast to the Tsai-Hill criterion, Tsai-Wu contains also some linear terms.

= F1 1 + F2 2 + F11 21 + 2F12 1 2 + F22 22 + F66 212

(3.105)

Also here, the failure condition is given by equation (2.47). The quadratic parts
Q and the linear parts L can be separated.

= Q + L

(3.106)

has to be suppleFormulating the matrix expression, the strength matrix S


TW
.
The
matrix
formulation
is
given
by
mented by a strength vector s
TW
T
Q = 12
STW 12

(3.107)

L = s
TW 12

(3.108)

and

3.6 Strength Sensitivities

81

The strength matrix S


and the strength vector s
contain the strength
TW
TW
parameters F i j and F i in stress space which are specified in 2.3.2.

F11
F12
=
S
TW
0

s
TW = F1

F12
F22
0
F2

0
0
F66

(3.109)

(3.110)

The transformation to the strain-based formulation is done analogical to the


derivation for the Tsai-Hill criterion using equation (3.98). The strength
ratio R can be found by extracting the positive root of the failure condition

Q R 2 + L R = 1

(3.111)

and yields
R=

L +


+

4Q + 2L

2Q

1
FI

(3.112)

The failure index sensitivities of the j th layer of the i th element using the TsaiWu criterion are given by
dF I i, j

dt

1 dR i, j
R 2i, j dt

(3.113)

with

dR i, j
dt

dt

L i, j +


+

4Q i, j + 2L

2Q i, j

i, j

(3.114)

For simplification, the evaluated expression is not written explicitly. The


derivatives of the quadratic and the linear part of the failure criterion are
d Q i, j

dt
d L i, j

= 2 T
x y,i STW j

= sTW j

d x y,i

d x y,i
dt

(3.115)

.
(3.116)
dt
dt
The strain derivatives are evaluated analogous to the Tsai-Hill formulation
and are given by equation (3.104). The failure index sensitivities using the

82

Sensitivities

y
x

(a)

(b) = 0

(c) = 45

(d) = 45

(e) = 90

Figure 3.13: Displacement field (a) and PSF-sensitivities (b-e) of a biaxially


loaded notched plate

Hoffman criterion [59] are derived identically. The only difference is the interaction term F12 which is given in 2.3.2.
A sample sensitivity field of the introduced PSF using the Tsai-Wu criterion is shown in Figure 3.13. It is a section of a plate with a circular hole
loaded with biaxial tension. The corresponding displacement field is displayed
in Figure 3.13(a). Again, the sensitivities are shown for orientation angles
(0/45/-45/90). They indicate that the strength can be increased maximally if
fibers are added in circumferential direction which is absolutely plausible. The
fields can be transferred into each other by performing a stepwise rotation of
45 . Differences between the fields arise from the irregular mesh.

3.7 Hybrid Evaluation

83

3.7 Hybrid Evaluation


As mentioned in the introduction of this Chapter, the finite element framework
must be accessible in order to be able to determine the analytical sensitivities.
This is not the case for commercial software wherefore the framework has to be
implemented basically by the user. It is practicable as long as only a small set
of element types are used. If the considered model contains a large number of
element types (e.g. shells, solids, beams, masses, rigid body elements etc.), the
implementation effort becomes enormous or even impossible. However, this
problem can be mitigated by taking advantage of a hybrid evaluation.
The design variables considered above are the layer thicknesses of layered
shell elements only. Thus, the derivatives of element stiffness- and massmatrices are needed only for the considered shell elements since all other
parts of the stiffness matrix are independent on the design variables. Subsequently, element stiffness- and mass-matrix, or its derivatives respectively,
are derived independently and implemented in a separate programming environment. All the parts that are dependent on the overall system matrices,
such as displacements, eigenvectors and eigenvalues or the system matrices
itself are determined with the commercial software which is in this case NASTRAN. Even if NASTRAN is used here, the concept can basically be combined
with any other commercial finite element software. Consider the sensitivity
NASTRAN

?

d n

dt

Tn

?
dK
dt

?
dM
dt

6T Mn 6
n

MATLAB

Figure 3.14: Hybrid evaluation scheme for eigenvalue sensitivities


equation for a dynamic eigenvalue problem which is given by equation (3.53).
Stiffness- and mass-matrix derivatives are evaluated in a separated environment which is here MATLAB. The eigenvalue and the eigenvector are result
of a NASTRAN-run and can be accessed easily. The mass matrix is not needed
explicitly since the generalized mass T M which is a scalar value provided
by NASTRAN. It is usually one for the dynamic case but not necessarily for the
buckling eigenvalue problem. NASTRAN provides the possibility of exporting
the global stiffness matrix K which is required if the displacement sensitivities
are needed (see equation 3.44). In order to obtain exact results evaluation the

84

Sensitivities

NASTRAN

du x
dt
MATLAB

?
=

1
K11

dr c
dt

dK11
dt

?
ux

Figure 3.15: Hybrid evaluation scheme for nodal displacement sensitivities

sensitivity equations, the formulation of the layered shell elements should be


equivalent to the formulation of the same element in the applied commercial
software. The convention of the degrees of freedom, the integration procedures
and the element shape function must be known therefore. The commercial
software providers do usually not publish these informations explicitly. For the
investigations within this thesis, a bilinear four-noded and a triangular threenoded element have been implemented which are represented in the commercial framework with the NASTRAN CQUAD4- and the CTRIQ3-element. These
simple elements are state-of-the-art and described in a couple of publications
[104, 105, 107, 106].
Using the proposed hybrid evaluation, the sensitivities for layered shell
elements can be evaluated in a finite element model with arbitrary complexity without knowledge of the remaining element types. Due to the analytical
derivation, no numerical error results as it is the case for a numerical approximation. Additionally, the danger of a wrong implementation of the finite element framework can be reduced since global results arise from the commercial
program.

3.8 Smoothing
The application of finite elements holds the problem that strong gradients in
the strain field may occur. The gradients are a result of the linear theory underlying the FEM-theory used here. They can usually be observed in regions near
load introduction points or structural inhomogeneities, e.g. sharp edges or material stiffness changes. Neighboring elements of loads or inhomogeneities may
be stressed several orders of magnitude higher than surrounding elements.
The same phenomenon can be observed for sensitivity fields. The sensitivities are connected to the loading of the finite elements wherefore the difference
between adjacent element may be high. This is problematic for an optimization
since the linear modeling overestimates the sensitivity of a few single design

3.8 Smoothing

85

variables to the objective function. Consequently, suboptimal or wrong solutions are result of the optimization process. To prevent the problem of strong
gradients, a smoothing of the sensitivity field is performed. The idea is inspired by the smoothing of pixels in image processing which is known as image
blurring.
Consider a vector g containing the sensitivities of all n elements of a layer.
g = [ g 1 , g 2 ,..., g n ]T

(3.117)

In order to smoothen this vector, the connectivity of the elements must be


known. This is caught by a table containing the information which nodes
belong to which elements. The table lists the nodes in the columns and the
elements in the rows.

n1

1
0

n2

for
for

nl

ni e j
ni e j

e1
e1

..
.

(3.118)

en

A value of 1 indicates that the node is part of the corresponding element. If


there is no relation between element and node, zero is entered. Multiplying
the affiliation table with its transpose yields an element connectivity table
which obviously expresses the connectivity between the elements.

= T

(3.119)

The entries contain the number of common nodes between the elements. A matrix multiplication of the element connectivity table with the sensitivity vector
yields a smoothened vector. However, the multiplication would also cause an
unwanted dilation of the vector. Thus, the connectivity table is normalized
such that the sum of each row is one. The normalized connectivity table is
then denoted by . The smoothing can be increased by raising the connectivity
table to the power of p. A value p of zero is equivalent to no smoothing
gsm = g

(3.120)

The smoothing process is explained schematically on a finite element model


consisting of six four-noded elements where the nodes are numbered from 1 to
12 and the elements from I to V I (see Figure 3.16). The corresponding affilia-

86

Sensitivities

10
IV

VI

5
I

7
II

12

11

III

Figure 3.16: Element cluster

tion table is given in equation (3.121).

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

0 0 1 1 0 0 1 1

0 0 0 0 1 1 0 0

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

10

0
0
0
1
0
0

11

0
0
0
1
1
0

0
0
0
0
1
1

12

0
0
0
0
0
1

I
II
III

(3.121)

IV
V
VI

As explained above, the ones indicate the nodes that belong to the element.
Thus, each row contains four entries with one (because of the four-noded elements) while the rest is zero. The connectivity table results of the multiplication with its transpose.
I

= T =

II

4
2
0
2
1
0

2
4
2
1
2
1

III

IV

VI

0
2
4
0
1
2

2
1
0
4
2
0

1
2
1
2
4
2

0
1
2
0
2
4

I
II
III

(3.122)

IV
V
VI

Diagonal entries express the number of common nodes of each element with
itself which is consequently the number of nodes of the element. Alternatively,
elements can be connected with two or one nodes, or they can be not connected
(only valid for the considered example). Naturally, this matrix is always symmetric because the connection of the nodes is two-way. The gradient smoothing

3.8 Smoothing

87

is illustrated on this example with a sensitivity vector of



g = 1

20

T
1

(3.123)

Thus, the sensitivity value of the second element is much higher than of the
others. The non-smoothened sensitivity is shown in Figure 3.17(a) where the
smoothing factor p is set to zero. Figure 3.17(b) shows the sensitivity field with
a smoothing factor of 1. The value of the second element has been decreased
while raising the values of the other elements. The values of the elements that
are connected to element II with two nodes have obviously a higher raise. The
smoothing with a factor of 2 is illustrated in Figure 3.17(c).
10

(a) p=0

(b) p=1

(c) p=2

Figure 3.17: Example of the smoothening of a sensitivity field

88

Sensitivities

Chapter 4

Automated Design Process


This chapter explains the automated design process for the generation
of local laminate reinforcement layers, which is later referred to as
Ghost Layer Method (GLM). The Ghost Layer Concept, which is an essential
part of the patch generation process, is defined and explained with an illustrative example. Consequently, the parameterization or the model is explained
and the adapting process with the different strategies is introduced. Finally,
model requirements and process limitations are discussed.

4.1 Ghost Layer Concept


The Ghost Layer Concept is an essential part of the design process for local
laminate reinforcements. A ghost layer carries information on material properties and material orientation but its thickness is specified to zero. It can be
understood as a virtual layer which is existent in the computer model. Assuming that the thickness of the k th layer is zero, the following equation must be
valid.
j

k =1

tk

j
1
k =1

t k = 0 for t j = 0

(4.1)

Entering that expression into equations for the stiffness matrix (3.9-3.12), the
contribution of the k th layer vanishes. Thus, a ghost layer does affect neither
the stiffness nor the mass and has no influence on the structural behavior at
all. However, the sensitivities of a ghost layer with respect to a change of its
thickness are not equal to zero. They can be determined with the regular sensitivity equations that have been derived in Chapter 3. Considering the sensitiv-

90

Automated Design Process

ities of the membrane stiffness (3.16) or the transverse shear stiffness (3.17),
the thickness variable t l does not occur. This is also the case for bending
and coupling parts. The sensitivity there is dependent on the stiffness of the
overlying layers. Thus, the sensitivity value does not vanish even if the layer
thickness is specified to zero. This can be understood also intuitively. An infinitesimal change of the layer thickness, even if it is zero, does always change
the stiffness of the laminate. The same is valid for mass and geometric stiffness sensitivities. Ghost layers provide the possibility of obtaining information
where and how an existing laminated structure should be modified in order to
raise the objective function maximally. To do so, an arbitrary number of these
virtual layers with different materials or orientation angles can be placed at
the interfaces of a real laminate. The obtained sensitivities indicate areas and
layers that affect the objective function the most by applying an infinitesimal
change of thickness. Based on that information, the structure can be adapted
in order to improve the structural behavior. More explanations and applications of the Ghost Layer Concept are given in [144, 145].
The application of these ghost layers is illustrated on a simple example of a
vibrating plate. It is made of an isotropic material and clamped on three sides.
The lowest eigenmode of the plate is plotted in Figure 4.1. Goal of this exper-

Figure 4.1: Lowest eigenmode of three side clamped isotropic plate


iment is to locate regions where the structure should be reinforced in order to
raise the frequency. Four ghost layers of an unidirectional reinforced composite material with orientations of 0 , 45 , 45 and 90 are stacked to the plate.
This is done symmetrically but only one symmetric half of the laminate is used
for the illustrations. The orientation angles are predefined in order to match
the requirements of manufacturing adequate designs. However, more layers
and/or different orientation angles could be used. Taking the eigenvector and
the eigenvalue of the isotropic plate and the element matrix derivatives of the
ghost layer, the sensitivities can be evaluated using equation (3.55). The re-

4.1 Ghost Layer Concept

91

sults for each of the four orientations are shown in Figure 4.2. Red colored

45

-45

90

Figure 4.2: Ghost layer sensitivities for the three side clamped plate
regions indicate high and blue regions low sensitivities. The highest values
thus occur at the edges of the 0 -layer. Slightly increased values can be observed and at the long clamped side of the 90 -layer. Thus, the frequency is
increased maximally by adding material of the corresponding orientation angle in these regions. For this simple example, it is intuitively comprehensible
that such reinforcements may be ideal. For more complex structures, the sensitivity field cannot be captured intuitively and it even may have a surprising
topology. Ghost layers or just parts of them can be turned into a real layer by
assigning a real thickness to them.

92

Automated Design Process

4.2 Parameterization
Based on the sensitivities, the thickness of the element layers can be changed
to find improved designs regarding the objective function. The thickness can
basically be set to arbitrary values. However, an important goal of the method
is to provide designs that can be manufactured without the need of making
major modifications. The thickness of the laminate plies is specified by the
semi-finished material provided by the suppliers. Any designs containing layer
thickness values varying from these materials are infeasible for production.
To avoid the existence of such solutions, the layer thickness is restricted to
predefined values. This is solved by specifying the layer thickness t ctrl based
on the material that will be used for the intended structure. Additionally, a
binary thickness control array t ctrl is introduced which decides if a layer is
existing or not. A zero entry sets the thickness to zero, while a unit value sets
it to the predefined material thickness. The total thickness is calculated by the
element-by-element product of these two arrays.
t i = tTctrl,i tmat,i

(4.2)

Consequently, the thickness control array t ctrl serves as design variable


vector x.
The parameterization scheme of the model is illustrated in

t1

t2

t3

t4

t5

t6

tm

tn

tmat

t ctrl

? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ?
t1

t2

t5

tm

tn

Figure 4.3: Parameterization scheme of the design process


Figure 4.3. The array tmat , which contains the thickness values of the physical material, is predefined and is not modified at any time. Thereby, a set of
thickness variables t i to t i+ x belong to a particular element. The thickness control array t ctrl maps the material thickness to the total thickness array t. The
existence of layers within all elements is thus coded with a simple bit-string.
Consequently, the design variable array has binary characteristics.

4.3 Search Direction

93

4.3 Search Direction


In order to find improved designs corresponding to a given criterion, a search
direction s is defined. The search direction vector can be understood as the
sensitivity field of the objective function f assuming that the objective is to be
maximized. In case of minimizations, the search direction has to be reversed
in order to point in the direction of steepest descent. The search direction is
here defined as a weighted function of the sensitivities derived.
s=


i

wi

d fi
dt

(4.3)

Using this definition, the objective function f is thus a weighted sum of its
sub-objective functions f i . Figure 4.4 illustrates the composition of a search
direction of the two sensitivity vectors of f 1 and f 2 for a 2-dimensional design
space with design variables x1 and x2 . The contours and the maximal point
for both functions are plotted. The sensitivity or the gradient vector, respectively, at the current design point (CDP) stands perpendicular to the contour
lines. The search direction is built by a vector addition of the two sensitivities. Keep in mind that the search direction is only exact for the current design
point. A design change, which would cause a relocation of the current design
point, results in a change of the search direction. The search direction vector
CDP
df2
dt

df1
dt

f2max

s
t2
t1

f1max

Figure 4.4: Search direction s as a weighted function of two gradient fields


has an entry for each design variable. It can be visualized equivalent to the
sensitivity field by coloring the element layers. Thus, the search direction field
can be interpreted as the superposition of the contributing sensitivity fields.

94

Automated Design Process

The construction of the search direction or, the combination of different sensitivities respectively, can be chosen arbitrarily by the user. In the following, a
couple of possible weighting functions which are suitable for structural design
are discussed. However, the variety of potential weighting functions is wide.
If the lowest eigenfrequency of a structure is to be maximized, the
search direction can simply be defined as the sensitivity field of the lowest
eigenvalue 1 .
s=

d 1
dt

(4.4)

Thereby, the objective function is the lowest eigenvalue 1 and the weighting
factor is set to 1.
Alternatively, not only the lowest but the n lowest eigenvalues may be considered. The goal is thereby to increase the frequency of all n eigenmodes. The
search direction is
s=

n

i =1

wi

d n
dt

(4.5)

and the objective function is now the weighted sum of the lowest n eigenvalues.
In order to weight the lower eigenvalues more, the weighting factor w i may be
2
set to the reciprocal of the square of the corresponding eigenvalue (w i =
).
i
However, other choices such as a homogeneous weighting might be more adequate in some cases.
Another optimization objective could be to make eigenmodes drift apart from
each other in order generate a resonance-free band. Doing that for the first
and the second mode, the objective function may be
s=

d 1 d 2
+
dt
dt

(4.6)

While the first mode is minimized, the second mode is maximized which results
in a drifting of the modes. A second method to free a frequency band is the
usage of the function
wi =

1
thres i

(4.7)

The threshold eigenvalue thres splits the frequency band into a lower section,
where frequencies are minimized, and a upper section, where frequencies are
maximized. Eigenvalues being close to the threshold are weighted more in order to push them away. The corresponding weighting function characteristics
is shown in Figure 4.5. The same weighting functions may be used for buckling

4.3 Search Direction

95

thres

Figure 4.5: -weighting factor function


problems.
For strength or stiffness problems, a weighting function may be helpful to consider different loadcases. The sensitivities calculated in Chapter 3 are dependent on a single load vector r only. In most case, a structure is however loaded
differently during a lifetime cycle. The search direction for stiffness may be
defined as
ri

n


dW
wi
s=
dt
i =1

(4.8)

and for strength as


s=

n

i =1

wi

i
d f PSF

dt

(4.9)

The weighting factor sets priority to loadcases that are more important. Also
a generation of multi-criteria sensitivities is feasible. For example, the lowest
eigenfrequency of a structure may be increased while simultaneously stress
peaks occurring for a specific loadcase are reduced. Such a search direction
may look like
s = w1

d 1
d f PSF
+ w f PSF
dt
dt

(4.10)

The choice of the weighting factors w1 and w f PSF is not trivial since the units
are different. The units can be eliminated by dividing the sensitivities with
the value of the corresponding objective function which changes the search
direction above to
s=

w1 d 1
w2 d f PSF
+
1 dt
f PSF dt

(4.11)

It will make sense to perform a test-run and check whether the structural
response is changing in an appropriate way. Design examples for the search
directions discussed above are shown in Chapter 6.

96

Automated Design Process

4.3.1 Search Direction as Design Support Information


The automated adaption process based on the search direction is discussed in
the next Section. However, the information of the search direction may already
be of value when designing a structure manually. Structural design is today
still often carried out in a trial-and-error procedure. Engineers try to improve
the structural design by using the information of displacement, stress, strain,
element strain energy, eigenmodes and more, and their experience and intuition. If designing laminate structures, the intuition may reach a limit even
if the complexity of the geometry is low. The information of a search direction, which provides layer-wise information on the characteristics of a composed global objective function, may be very helpful. Especially, if the objective
function is a superposition of more than one sensitivity field, this information might be valuable. High values of the search direction indicate where the
structure must be reinforced and vice versa. Based on that information, the
structural behavior can manually be improved more specifically and more efficiently. Figure 4.6 illustrates the complexity of designing a structure if the

(a) Mode 2

(b) Mode 3

45

-45

90

(c) Sensitivities Mode 2

(d) Sensitivities Mode 3

(e) Sensitivity Superposition

Figure 4.6: Superposition of the sensitivities of mode 2 and 3 with w2 = 1 and


w3 = +1
search direction is a superposition of two sensitivities. Thereby, the modes 2
and 3 of the three side clamped plate are considered. Corresponding to the

4.3 Search Direction

97

weighting methods introduced above, the frequency of the second mode has
to be minimized while the third frequency is maximized. Thus the weighting
factors are set to w2 = 1 and w3 = +1. Figure 4.6(c) and 4.6(d) show the sensitivity field of the second and the third mode in Figures 4.6(a) and 4.6(b). The
superposed search direction is shown in Figure 4.6(e). An optimal design of
the particular modes may be found with experience and intuition, even if the
potential optimal reinforcements are not obvious without the sensitivity field.
However, finding good designs without the search direction information is almost impossible.
Taking advantage of the search direction field in Figure 4.6(e), a simple reinforced solution could look as shown in Figure 4.7. Corresponding to Table 4.1,

45

-45

90

Figure 4.7: Potential solution of manually generated reinforcements based on


the search direction field
the second frequency has been increased by 6.32%. A frequency drop can here
not be reached since only reinforcement and no material reduction is allowed.
However, the third frequency was raised by 35.39% which is significantly more
than for the second mode. Better solutions can be found if the model is modified
stepwise as done in the following proposed adaption process.

basic design
reinforced design

f2 [Hz]
428.81
455.93

(+6.32%)

f3 [Hz]
629.36
852.12

(+35.39%)

Table 4.1: Numerical results for the manually reinforced plate

98

Automated Design Process

4.4 Adaption Process


The adaption process aims for solutions with an improved structural behavior compared to the reference design. Based on the search direction s arising
from the sensitivity analysis, the binary design variable array t ctrl is modified.
Basically, the optimization problem can be formulated as a topology optimization problem. The objective function f has to be maximized referring to the
underlying finite element problem.
Objective: max

f (t ctrl )

Subject to:

t ctrl

g< g

t ctrl {0,1}
The optimization continues as long as the termination criterion is not fulfilled. This is expressed with g < g. Thereby, g must be lower than a given
parameter g. This could for example be the model mass which may not exceed
a predefined value. Any other termination condition may be used, e.g. number of iteration loops, frequency, displacement, buckling load factor, margin
of safety and many more. However, it is not guaranteed that the termination criterion can be fulfilled during the adaption process. The solution space
is restricted to the provided set of ghost layers and not any solution can be
achieved.
Basically, there exists no gradient-based optimization algorithm to
solve binary problems. Common gradient-based optimization algorithms,
like Cauchys method of steepest descent [22] or the conjugated gradient
method [37] cannot be employed without modifications due to the binary characteristics. Bendse and Kikuchi [15, 14] avoid this problem by introducing a
density function which transforms the binary formulation of a topology problem into a continuous formulation. Their density function is continuous and
bounded between 0 and 1. To obtain more realistic solutions, they are pushed
towards the boundaries before terminating the optimization. This approach
has not been used here. The method holds the risk of getting structures with
scattered distribution of the layers, similar to a checkerboard. From the mechanical point of view, the checkerboard pattern could be interpreted as layers with reduced thickness whose stiffness is smeared over a certain area.
However, such solutions are inappropriate for manufacturing processes. Additionally, continuous design variables may cause problems using the hybrid
evaluation technique proposed in Section 3.7. The continuous variables will
unavoidably lead to a different layup in each element of the model. Thus, each

4.4 Adaption Process

99

element needs its own laminate property card (PCOMP). This must be handled
in the MATLAB framework which has turned out to be computationally expensive.
Considering other classes of potential optimization algorithms to find solutions
of the above defined optimization problem, the stochastic algorithms [159] and
especially the Genetic Algorithms (GA)[60] have to be mentioned. Numerous
variations of stochastic algorithms exist that are able to solve the problem.
Even if some more sophisticated hybrid algorithms take advantage of gradient information, conventional stochastic ones do not make use of sensitivities
since the idea is based on random processes. However, the sensitivities provide
important information how an optimal design may look like. This information
should be employed in any cases. GA work only efficiently when evaluations of
the objective functions are executed in parallel what calls for a large number
of finite element software licenses. This is usually not given in an industrial
environment since licenses are expensive. Due to these facts, the usage of
stochastic algorithms has been avoided.
The method for the design of the locally reinforced structures, called Ghost
Layer Method (GLM), is far more simple and pragmatic. A flowchart of the
method is given in Figure 4.8. In the first step, the initial model is read and
stored in a database which is here MATLAB-based. With this information, the
evaluation of the sensitivities is carried out. Based on the obtained search
direction s, the thickness control array t ctrl is modified. The process is continued until the termination criterion stops the process.
The adaption process, which is beside the evaluation of the finite element
model and the sensitivities another important part of the design process, is
discussed here in detail. The search direction s points into the direction of
the best design improvements. Due to the binary characteristics of the design
variables, the design can however not be modified according to that direction.
Figure 4.9 illustrates the search direction and the vectors of the three potential
design changes (colored green). The thickness variables t1 or t2 , or both, can
be set to a value of 1 in order to find an improved design. However, not only the
direction but also the step size should be considered when modifying the design
variable array. If both variables t
1 and t 2 are set to 1 simultaneously, the step
size becomes bigger (in this case 2) than if only one variable is changed. If m
entries of the thickness control array t ctrl are adapted simultaneously, the step

size is m. The more design variables are changed within one iteration loop,
the bigger the distance to the new design point in the search space. However,
the search direction is a linear extrapolation of the respective objective function gradient at the current design point. Each change of a design variable has
an influence on the structural behavior of the model and on the search direc-

100

Automated Design Process

Initial Design
input-file
Reading

Storage
model
tctrl

Evaluation

Writer
input-file
FE-Evaluation

System Matrix
Sensitivities

Adaption
Process

output-file
Reader
f,u,,,K

dK, dM, dK
dt dt dt
Sensitivity Evaluation

NASTRAN

Termination

tctrl

Final Design
Figure 4.8: Flowchart of the automated design process

4.4 Adaption Process

101

t2
1

s
t1

0
0

Figure 4.9: Linear assumption of the sensitivities

tion thus. If many design variables are changed simultaneously, the error due
to the linear assumption of the sensitivities will be bigger. To keep the error of
predicting the objective function characteristics with the search direction low,
a new evaluation of the gradient field should be performed after a modification
of one single design variable. This is possible, but many iteration loops are
needed until a final solution is obtained. Therefore, it makes sense to change a
group of design variables in one step before reevaluating the search direction.
However, it is not defined properly how many design variables are modified in
one iteration loop or before the search direction is reevaluated, respectively. A
sensitivity-threshold needs to be defined which decides whether a design variable is changed or not. The threshold can be chosen in many different ways.
A simple method is to set the threshold in percentage of the range between
minimal and maximal value occurring in the search direction. Consider again
the example of the three-edge clamped plate introduced above (see Figure 4.1).
The sensitivity values are plotted in Figure 4.2 and arranged in a histogram
(see Figure 4.10) which gives a visual impression of their distribution. The xaxis represents the range of the sensitivities and the y-axis gives the number
of elements which are within the sub-ranges. Each of the 20 bars quantifies
the number of sensitivities lying in the sub-range of 5% in a logarithmic scale.
Obviously, the bulk of the sensitivities are rather low. There exist a few element layers whose sensitivity values are high compared to the average value.
According to Figure 4.2, they can be located at the edges of the 0 -ghost layer.
Assuming that the lowest frequency is to be maximized with minimal mass addition, the plate has to be reinforced with a 0 -layer in these areas. A threshold

102

Automated Design Process

10

Counts

10

10

10

10

Sensitivity Range

Figure 4.10: Histogramm of the sensitivities of the three side clamped plate

has to be defined that decides which element layers are turned to real layers
by setting the thickness control variable to 1. Setting this threshold to, for example 5%, all elements that contribute to the very right bar of Figure 4.10 are
modified within one iteration loop. The resulting threshold is indicated with
the green line. Another visualization of the percentage-threshold is provided in
Figure 4.11. On the left side, the sensitivity fields of Figure 4.2 are plotted in
the same axis. Additional to the color map, the z-axis indicates the sensitivity
value of the corresponding element layer. The red colored peaks are part of the
0 -layer and the green colored bulge on the back edge belong to the 90 -layer.
The threshold is visualized with the white rectangle. In Subfigure 4.11(a), the
threshold is set to 10% between highest and lowest value. Obviously, only the
two corner elements of the 0 -layer exceed that value. These two elements can
also be identified by the very right bar in the histogram in Figure 4.10. The
corresponding plot on the right side (Subfigure 4.11(b)) shows the same picture from the top whereas all parts below the threshold surface are hidden.
This plot can be understood as the reinforcement layer that develops when the
threshold is set to 10%. To show the influence of the threshold choice on the
reinforcement design, the same plots are shown for a threshold of 50% (Subfigures 4.11(c) and 4.11(d)) and 80% (Subfigures 4.11(e) and 4.11(f)). The higher
the threshold, the bigger the reinforcement area generated within one iteration
loop. For the 80%-threshold, all four potential reinforcement layers are growing. As discussed, the step-size is thus large which may reduce the quality of
the solution. Considering eigenfrequency and buckling problems, there might
be the additional problem of mode switching. This means that the eigenvalue of
the originally second eigenmode drops below the eigenvalue of the first mode as
a consequence of the reinforcements. This might change the objective function

4.4 Adaption Process

103

(a) Sensitivity topology with threshold 10 %

(b) Reinforcement with threshold 10 %

(c) Sensitivity topology with threshold 50 %

(d) Reinforcement with threshold 50 %

(e) Sensitivity topology with threshold 80 %

(f) Reinforcement with threshold 80 %

Figure 4.11: Schematic illustration of the thickening process with the


percentage-threshold

104

Automated Design Process

and the search direction completely. Making only a small adaption within one
iteration loop, the method is more robust and flexible to react in case of a mode
switch. The method of generating reinforcements with percentage-threshold
can be used analogously for a reduction of material. There, the threshold is
set in order to reduce the thickness of element layers with low sensitivities.
Recalling the distribution of the sensitivities in Figure 4.10, it becomes clear
that the percentage-threshold must be chosen differently since there are many
element layers within the lowest range. Such a distribution has by the way
been noted for any investigated examples. It is difficult to find a threshold that
picks an adequate number of elements to be reinforced. Therefore, it might be
better to choose another strategy when performing material reduction.
Instead of choosing a set of element layers by percentage of the sensitivity
range, a fixed number of design variables could be changed within one iteration loop. Consequently, the number of steps, until a certain amount of mass is
added, can be predicted easily. This might be helpful to estimate the potential
improvements reached with the reinforcements. The number has to be chosen
by the user dependent on the number of design variables. The choice is thus
different for each model. A drawback might be that the relative difference between the sensitivities is not taken into account wherefore element layers with
totally different sensitivities are treated equally.

4.4.1 Limitations
The sensitivities derived in Chapter 3 express the influence of an infinitesimal
thickness change in one layer of one finite element. These values are changed
when the structural model is changed and they have to be reevaluated after a
modification of the model. To keep the calculation cost reasonable, a few element layer thicknesses are adapted within one iteration loop which however
causes an error. The error is small as long as the layers to be adapted belong
to different elements. In case that the thicknesses of layers of one element are
changed simultaneously, the error may become significant. It is well known
that the bending stiffness increases by the thickness in the power of three.
The sensitivities catch the stiffness change caused by the thickening itself and
caused by the fact that overlying layers are pushed outwards. If thickening
two layers of an element simultaneously, the sensitivity of the lower does not
contain the information that the upper layer is contributing additional stiffness to the laminate due to the parallel-axis theorem. On the other hand, the
sensitivity of the upper layer does not catch that it is pushed outwards while it
changes its thickness. If the relative position in the laminate stack is changed,
the linear assumption of the sensitivities may change significantly. It is there-

4.5 Modeling

105

fore recommended to only change one layer thickness per element within one
iteration loop. This is of course the one with the highest sensitivity value of
the element. If not doing so, the resulting reinforcements may be sub-optimal.
This limitation can be neglected for models that are only loaded in-plane since
the stacking position does not influence the membrane stiffness.

4.4.2

Comment

The presented thickening process might seem to be rather primitive since it is


not based on a sophisticated optimization strategy. It is inspired by the manual design process carried out by an engineer. The search direction is of high
value since it provides the information where the laminate needs to be adapted
in order improve the objective function or the structural behavior, respectively.
The adaptive process considering the search direction is appropriate to be done
by a computational system even if the work could also be done manually. The
process will usually not find the global optimum solution, except for simple
geometries and loadcases. However, finding global optimal solution is an important aspect of academic research, but is not one of the main objectives in
industry, especially in preliminary design. Moreover, proving that a solution
is optimal is usually impossible for complex optimization problems; otherwise,
the solution can easily be found. The goal of the design process in the industry
must be to find, with reasonable effort, solutions that have improved properties compared to an existing benchmark-design. The presented GLM has
turned out to be able to improve the design compared to the initial solution
significantly. Moreover, the obtained solutions are mostly producible without
major adjustments. The final solution can still be adapted manually in order to
match the manufacturing requirements better. It is possible that better solutions are obtained using more sophisticated optimization strategies. However,
the robustness of the process and the quality of the solutions does not indicate
rejection of the proposed method.

4.5 Modeling
4.5.1

Model Requirements

As mentioned in the Section discussing the hybrid evaluation 3.7, the tool developed within this thesis is based on MATLAB and NASTRAN. While the complete management, optimization and the evaluation of sensitivities is carried
out in MATLAB, the finite element run is made with NASTRAN. All the comments concerning model requirements are made with respect to these two soft-

106

Automated Design Process

ware products. However, the presented method can be implemented to any


other equivalent environments. The part of MATLAB can be assumed by any
programming or script language that provides the same basic commands. The
finite element run can be performed by any finite element software. Naturally,
the interface between the two programs would have to be organized differently.
The starting model must be available in the NASTRAN-input format,
more specifically in the small field format (see NASTRAN Quick Reference
Guide [117]). All elements that are available in NASTRAN can be used. However, the reinforcements can only be generated on layered shell elements and
here more specific for a simple four-noded quadrilateral element (CQUAD4) and
a three-noded triangular element (CTRIA3). More sophisticated elements, e.g.
a quadratic one, have to be implemented which has not been done in the framework of this thesis. Additionally, laminated parts of the model have to be defined with the PCOMP-property card. The material cards that are interpreted
for laminates are MAT1(isotropic), MAT2(anisotropic), MAT8(orthotropic). The
material orientation of each element has to be defined by an angle giving the
deviation of the element coordinate system. Concerning the elements above,
this is done with the THETA-argument which must be a float number. The
size of the model is limited by the required sensitivities which build the objective function. If displacement sensitivities are utilized, the processable model
size becomes moderate since the need of computational power and the physical
memory is increased.

4.5.2 Mesh Size


The chosen parameterization scheme causes the shape of the generated reinforcement layers to be discretized by the mesh of the finite element model.
Thus, the resulting reinforcements are dependent on the choice of the finite
element mesh size. In general, the contours or the reinforcements become
smoother if the element size is chosen to be small. The generated solutions
may be easier to convert into designs that are feasible for production. The
errors made by the simplifications are smaller wherefore the simulation represents the real physical behavior more accurately. However, it must be considered that the mesh size has an enormous impact on the calculation costs. The
finite element run, but also especially the evaluation of the sensitivities increases disproportionately high to the number of elements. Furthermore, the
needed physical memory may limit the number of elements. Before starting
the optimization, the user must choose an appropriate finite element mesh.
It has been observed that the obtained reinforcement layer shapes are not
strongly dependent on the choice of the mesh. To demonstrate the dependency,

4.5 Modeling

107

the three-edge clamped plate introduced above (see Figure 4.1) is optimized using different mesh sizes. The plate may be reinforced with fiber orientations
(0/45/-45/90) and the process is stopped when a certain mass is reached. The
fine mesh is generated by splitting the elements of the rough mesh into four
smaller elements (see Figure 4.13).
The characteristics of the process are

First Eigenfrequency [Hz]

600

500

400

300

200

100
0.052

rough
fine
0.054

0.056

0.058

0.06

0.062 0.064
Mass [kg]

0.066

0.068

0.07

Figure 4.12: Characteristics of the reinforcement generation process for different mesh size
shown in Figure 4.12. Obviously, the frequency calculated by using the fine
mesh is between 3.5 % and 6 % lower than the one obtained with the rough
mesh. The finite element model discretizes the fully continuous field into a
piecewise continuous field within each finite element. The resulting numerical error of the discretization can be reduced by using more elements. Thus,
models with more finite elements come closer to the exact solution. While the
initial solutions are equal in terms of laminate layup, the local laminate differs for the reinforced solutions due to the slightly different contours. Consequently, the error between the two meshes increases while reinforcements are
generated.
For the applied shell elements here, the shear locking effect may influence the
discrepancy between the two models as well. One of the basic assumptions of
the layered shell element is that the thickness is small compared to the other
dimensions. Even if this may be valid for the entire structure, the thickness
of a single element may be in the range of its other dimensions. Nevertheless, the single element has no knowledge about the surrounding elements.
Changing the in-plane dimensions of the shell element, the contribution of
bending stiffness and transverse shear stiffness does not change proportionally. Having thin elements, the contribution of the transverse shear stiffness is

108

Automated Design Process

(a) = 90 rough

(b) = 90 fine

(c) = 45 rough

(d) = 45 fine

(e) = 45 rough

(f) = 45 fine

(g) = 0 rough

(h) = 0 fine

Figure 4.13: Locally reinforced solution of a plate with different mesh size

4.5 Modeling

109

overestimated which is known as shear locking effect. This effect is prevented


by weighting the parts with the shear correction factor which is again dependent on the element dimensions (see equation (2.104)). Even if that error is
corrected well by commercial FEM-software, a slight error may remain. The
discrepancy between the rough and the fine model is acceptable at least for
preliminary design.
The reinforcement layers obtained for the finite element models with different mesh size are shown in Figure 4.13. Obviously, the contours of the two
models are almost identical. As expected, the contours of the fine model are
smoother which enables a more exact production of the designs. However, it
must also be considered that the calculation time of the fine model was higher.
Choosing the same settings, the processing of the fine model took approximately 5.5 times longer than the rough one. Not only the evaluation of the
model takes longer, but also more iteration loops are needed until the required
mass is reached. Of course, the calculation time can be reduced by choosing
different settings. Similarities for other models with different mesh sizes have
been observed as well. The user has to provide an appropriate finite element
mesh considering requirements and resources available. Dependent on the
calculation time, solutions obtained with the method can be double-checked by
using a different element mesh.

4.5.3

Physical Errors

The mechanics of layered shell elements which has been discussed in


Section 2.4.3 are based on the Classical Laminate Theory (CLT). The main assumption of the CLT is a plane stress state which implies that no stress components with out-of-plane components occur. This assumption is correct for thin
laminates with homogeneous thickness. Close to the edges of the laminate, the
assumptions do not longer hold. There, the in-plane stresses must vanish due
to the free surface. The static equilibrium can only be fulfilled when out-ofplane and interlaminar stresses counteract. These phenomena are known as
boundary effects or free edge effects. Similar effects can be noticed having an
inhomogeneity in the stiffness of the structure as it is the case at the edges of
local reinforcements. A stiffness inhomogeneity provokes stress peaks which
have to be considered when analyzing a structure in detail. However, layered
shell elements are not able to fully catch these effects due to the simplifications. The stress fields close to the laminate edges or close to edges of local
reinforcements have been investigated in various ways. A detailed overview
of investigations on free-edge effects is given by Mittelstedt and Becker [114].
While the stress field on free edges can be assumed to be two-dimensional,

110

Automated Design Process

the stress field becomes three-dimensional considering free corner effects. Extensive investigations have been made by Becker et al. [10] and Wigger and
Becker [179, 180, 181, 182] using both, analytical and numerical approaches.
In order to demonstrate the incomplete stress information near stiffness
inhomogeneities provided by shell elements, a simple comparison is presented
here. The considered laminate has a layup of (0)4 and a rectangular 0 reinforcement layer is attached at the top. The laminate is uniaxially loaded
(displacement controlled) and a cross section parallel to the loading path is
considered (see Figure 4.14 A-A). Using layered shell elements, the model pro-

y
x
Figure 4.14: Uniaxially loaded plate with local reinforcement layer
vides the stresses x , y and x y (all other components are assumed to be zero).
Figure 4.15(a) faces the stresses at the top basic layer arising from a shell
model (colored blue) and an equivalent model made with solid elements (colored red) over the distance of the considered cross section. At position x=0.5,
the reinforcement layer drops wherefore an inhomogeneity in the stiffness or
in the stress field, respectively, occurs. The stress x increases abruptly. While
the stress curve of shell model increases by leaps and bounds, the curve from
the solid model shows an extensive overshooting of the stresses. Similar overshooting can be noticed for the stresses y and x y . Consider that these two
components are very low if compared to x for the current loadcase. Analogously, Figure 4.15(b) shows the in-plane stresses in the reinforcement layers.
Also here, the curves of the different modeling manners do not closely match
the reinforcement layer drop. Qualitatively, the forces in the reinforcement
layer are transferred to the basic layer. Basically, it seems that the modeling with shell elements underestimates the stresses close to stiffness inhomogeneities.
However, also the ideal modeling with solid elements has some problems which
are discussed here briefly. Figure 4.16 shows the polished specimen photomicrographs of laminates which both contain a layer drop. They have been prepared within the Semesters thesis of B. Rentsch [140]. Figure 4.16(a) shows
the dropping of a middle layer which is embedded between two other layers.
Obviously, the fibers end rather abruptly. However, the entire space on the

4.5 Modeling

111

0.5

0.1

xy [MPa]

400

200
0

xy

10

y [MPa]

x [MPa]

600

5
0
5
0

AA

0.5

0.1
0

AA

0.5

AA

(a) In-plane stresses in top basic layer

100
0.5
AA

xy [MPa]

200

0
0

xy

y [MPa]

x [MPa]

300

0.1

10
0

0.5

0.2
0

AA

0.5

AA

(b) In-plane stresses in reinforcement layer

Figure 4.15: Stress distribution for shell modeling (blue) and solid modeling
(red)

right side is filled with matrix material. This material forms a bevel which is
roughly indicated with the red lines. A similar situation can be observed for
an outer layer shown in Figure 4.16(b). The polished specimen is embedded in
epoxy which is almost identical to the laminate matrix material. Thus, the contrast between the two components is low and the layer boundaries have been
traced with red lines.
Even if the matrix material is soft compared to the fibers, the emerging bevel
may have a significant impact on the stress distribution. It helps to pass the
forces over the reinforcement layer. Alternatively, the bevel can be understood
as component that reduces the stiffness discontinuity which leads to a reduction of the stress concentrations. If modeling a layer drop with solid elements,
this should be considered in order to get realistic results. Figure 4.17(a) illustrates an ideal modeling as it has been made for the results presented above.
The bevel can simply be modeled adding a wedge element of the matrix material which is sketched in Figure 4.17(b). The results of the different modeling
methods are compared in Figure 4.18. The overestimation of the stresses ob-

112

Automated Design Process

(a) internal drop

(b) external drop

Figure 4.16: Photomicrographs of two layer drops

(a) ideal modeling


A

(b) modeling with bevel

Figure 4.17: Layer drop modeling

tained by the ideal model (red curve) is evident for the normal stresses x , y
and z . The stresses occurring when modeling with bevel (green curve) are significantly lower. In the far field, the two modeling methods of course provide

4.5 Modeling

113

0.5

4
2
0
2
0

AA

xy

10
0

yz

0.5
AA

zx

0.5
0

0.5
0

10

zx [MPa]

0.5
AA

yz [MPa]

xy [MPa]

0.5

AA

0.1

0.1
0

10

z [MPa]

400

200
0

y [MPa]

x [MPa]

600

0.5
AA

0
10
20
30
0

0.5

AA

Figure 4.18: Stress distribution for ideal modeling (red) and modeling with
bevel (green)
identical results.
The underestimation of the stresses using shell elements is obviously not as
high as expected when considering Figure 4.15. Nevertheless, the stress peaks
close to the layer drop still exist and have to be considered in the analysis. Performing a preliminary design, which the proposed method has been developed
for, the results of shell elements may however be adequate. Keep in mind that
the errors have been discussed on the example of a simple unidirectional laminate. Having a more complicated layup, the stress distribution may be more
complex. However, the stress peaks will occur for any cases.

114

Automated Design Process

Chapter 5

Verification
Within this chapter, the proposed Ghost Layer Method (GLM) for the design of
locally reinforced structures is verified by comparing its predictions of optimum
design solutions with those of established methods. The latter can solve a
limited range of suitable sample problems which are recalled and used in the
following sections.

5.1 Problem Formulation


5.1.1

Finite Element Model

The finite element model for the verification of the GLM consists of 100
quadratic four-noded shell elements which are stringed together as shown in
Figure 5.1. Each has a side length of 0.01 so that the entire length of the

Figure 5.1: Finite element model for the comparative study


model is 1. The loads and boundary conditions differ for the several load-

116

Verification

cases. In order to be able to find optimal thickness distributions analytically,


isotropic behavior is assumed. However, the basic concepts are the same as for
anisotropic materials. All parameters, variables and calculations are dimensionless wherefore no units are given for this study. The model parameters of
the beam are listed in Table 5.1. Thereby L denotes the length, b the width,
h the core thickness and t pl y the thickness of a single ply. E pl y and E core are
dummy Youngs moduli for the ply and the core material. The densities are
denoted analogically with pl y and core . The Youngs modulus of the core maGeometry

1
0.01

b
t pl y

0.01
1E-4

E pl y

1E12

pl y

1000

E core

1E9

core

1E-9

L
h
Ply / Facesheet
Core

Table 5.1: Dimensionless model parameters


terial is much smaller than that of the ply material in order to ensure that the
stiffness is governed by the facesheet which is the part to be optimized. The
density of the core material is set to a very low number but greater than zero
to prevent numerical problems.

5.1.2 Reference Solutions


For the verification, investigations on beam and bar structures are made. The
general differential equation for the behavior of a thin beam according to the
Euler-Bernoulli beam theory is


2
x2

EI z (x)

2 v(x, t)
x2


= (x)

2 v(x, t)
t2

+ q(x)

(5.1)

where v(x, t) are the lateral deflections dependent on the position x and the
time t. The general differential equation for a bar can be written as

E A(x)

u(x, t)
x


= (x)

2 u(x, t)
t2

+ p(x)

(5.2)

where the unknown axial displacements are denoted with u(x, t). In both equations, (x) is the specific structural weight or the mass per unit length. q(x)
and p(x) are the forces per unit length in lateral and axial direction. The cross

5.1 Problem Formulation

117

section area and the area moment of inertia are denoted with A(x) and I z (x) respectively. Here, only rectangular cross sections are considered wherefore they
can be expressed as a function of the thickness t(x) and the constant width b.

(5.3)

A(x) = bt(x)

I z (x) =

bt(x)
12

(5.4)

Searching for beam or bar structures with optimal thickness distributions t(x),
A(x), I z (x) and (x) become a function of location x. In this case, it is generally not possible to find an analytical solution. However, there exist some approaches for specific loadcases and geometries. Prager and Taylor [134] formulated so-called optimality criteria for different problems of optimal structural
design aiming for minimal weight. They claim that the specific stiffness s(x),
which is in case of a bar the product of Youngs modulus E and cross section
area A(x), and in case of a bending beam the product of Youngs modulus E
and the area moment of inertia I z (x), must be a linear function of the specific
structural weight (x). Considering a bar structure, this requirement is always
fulfilled since
s(x)bar = E A(x) A(x)bar

(5.5)

(x) bar = A(x) A(x) bar

(5.6)

and

However, for beam structures this relation is not given in general. Only if
assuming sandwich structures with rectangular cross sections and the core
thickness h much bigger than the facesheet thickness t (h  t), the condition
is approximately valid. Using the parallel axis theorem, the area moment of
inertia I z (x) and consequently the specific stiffness s(x) become a linear function of the cross section area A(x) of the facesheets.
t
(x) =
shbeam

1
Eh2 A(x) f acesheet A(x) f acesheet
2

(5.7)

Consequently, the specific structural weight becomes proportional to the specific stiffness as well.
beam (x) A(x) f acesheet

(5.8)

The analytical solutions of Prager and Taylor are only feasible to verify the
method for membrane loads. If considering bending beams, the analytical solutions assume a sandwich layup in which the governing stresses arise from
the membrane loads as well.

118

Verification

A couple of publications [119, 75, 121, 76, 77] address the problem of finding
beams with variable thickness distribution without restrictions to sandwiches.
However, explicit solutions are only found by taking advantage of numerical
methods. In order to find general beam solutions with variable thickness distribution for bending loadcases, a finite element model with beam elements
has been implemented. Stiffness, mass and stress stiffness matrices have been
taken from Cook et al. [28] and are given in Appendix C. According to the shell
finite element model for the verification, the reference beam model consists of
100 elements. The thickness t of each beam element is taken as design variable. A pre-implemented optimization algorithm of MATLAB [112], namely an
interior-point algorithm, has been used to find the optimal thickness distribution. A more detailed explanation of the algorithm is given in [19, 20, 176].
In the following sections, the obtained solutions with the GLM are verified comparing them to analytically and numerically derived solutions. For
sandwich beams and bars, the analytical solutions of Prager and Taylor are
employed. This approach provides solutions that are guaranteed to be globally
optimal. Loadcases with non-negligible bending loads are compared to the numerically determined solution. It is assumed that these solutions are global
optima as well but this can however not be proved.

5.2 Minimal Compliance


Considering a static loadcase, the Euler-Bernoulli equation simplifies to
EI z (x)

2 v
x2

= M b (x)

(5.9)

The cross section and the area moment of inertia I z (x) respectively are assumed to be functions of the location x. Prager and Taylor [134] formulate
an optimality criterion for the elastic design for maximum stiffness of sandwich structures which is reformulated in [133]. The specific stiffness s(x) for a
sandwich beam is given by equation (5.7). The specific structural weight (x)
becomes a linear function of s(x) wherefore minimizing the weight means minimizing the integral of s(x).

(5.10)
m = C s(x)dx

With this assumption, it is proved that the weight of a sandwich beam structure for a given deflection is minimal if
M b (x) M b (x)
(EI z (x))2

= const.

(5.11)

5.2 Minimal Compliance

119

Thereby, M b denotes the bending moment and M b the bending moment caused
by a dummy unit load. This optimality condition can also be used to find a
design with a prescribed weight and a minimal deflection. The mass, or the
volume respectively, is taken as side constraint and assumed to be constant.
The total volume of both facesheets is given by
V0 = 2

L
0

[b t(x)] dx = 2L b t0

(5.12)

where t0 is the facesheet thickness (which is set here to 5 t pl y ) of a beam


with constant section and equal weight. The thickness distribution t(x) for
maximum stiffness must fulfill equations (5.11) and (5.12).

5.2.1

Loadcase 1: Single load

Analytical Solution
Consider a sandwich cantilever beam which is clamped on the left side and
loaded with a single force F on the right side as shown in Figure 5.2. The
t
F

E , b, L

Figure 5.2: Sandwich cantilever beam with single load


core thickness h is assumed to be constant and much bigger (h  t) than
the variable facesheet thickness t. For the given loadcase above, the bending
moment M b is
M b (x) = F (L x)

(5.13)

In order to be able to apply the optimality criterion (5.11), a point x where the
deflection v(x) is considered must be defined. Here, the deflection at the right
end of the beam x = L is taken. This is done by applying a bending moment
caused by a dummy unit force at x = L which is
M b (x) = (L x)

(5.14)

120

Verification

Compliance is usually expressed with the potential of the external forces W


which is defined as
L
(5.15)
q(x)v(x)dx
W=
0

where q(x) is the distributed load. Because a single load is applied here, the
potential of external forces is
(5.16)

W = Fv(L)

wherefore a minimization of W and the deflection v(L) is equivalent. Using


equations (5.13) and (5.14), the optimality criterion forms to
M b (x) M b (x)
(EI z (x))2

F (L x)2
(E I z (x))2

=C

(5.17)

where C is a constant parameter. Solving this equation for the area moment of
inertia I z (x) and using the parallel axis theorem leads to


I z (x) =

F
C

1
2

 2
Lx
h
=2
bt(x)
E
2

(5.18)

Rearranging this equation results in an expression for the optimal facesheet


thickness distribution which is obviously linear in the variable x.


F
t(x) =
C

1
2

 
Lx 1 2 2
E 2b h

(5.19)

However, the expression is still dependent on the constant parameter C which


can be determined using the volume constancy condition (5.12).
  !
L    1
F 2 Lx 1 2 2
b
dx
V0 = 2
C
E 2b h
0
 1
  
F 2 1 1 2 2 L
= 2b
(L x) dx
C E 2b h
0
 1  
2 F 2 L 2
=
E C
h

C=

2
EV0

2

 4
L
h

(5.20)
(5.21)
(5.22)

(5.23)

5.2 Minimal Compliance

121

The optimal thickness distribution is found by combining equations (5.19) and


(5.23).
t(x) =


V0 (L x)
x
=
2t
1

0
L
b L2

(5.24)

The comparison between the analytical and the numerical results obtained
with the GLM is illustrated in Figure 5.3. The green layers represent the sand-

0.2

0.4

0.6

0.8

Figure 5.3: Thickness distributions for a single load (analytical)


wich core which has not been part of the optimization. The facesheet layers
which are result of the design process are colored blue. Consider that the beam
length, sandwich layers and facesheets layers are not plotted in true scale (and
neither for the following examples). The analytical thickness distribution is indicated with the red line. The layer distribution using the GLM matches the
analytical solution fairly good and the linear shape is clearly recognizable. Differences arise from the discretization within the finite element model.
The deflection v of the analytical design is determined solving the static
Bernoulli-Euler differential equation (5.9). Taking advantage of the boundary
conditions of the left hand clamping (v (0) = 0 and v(0) = 0), the deflection can
be written as
v(x) =

FLx2
2bEh2 t0

(5.25)

The maximal deflection vmax occurs on the right side x = L wherefore

vmax = v(L) =

FL3
2bEh2 t0

(5.26)

The numerical values are listed in Table 5.2. The stiffness of the GLM solution is slightly higher then the analytical solution but still in an acceptable

122

Verification

GLM

vmax

0.914 103

vmax

1.000 103

(+9.47%)

vmax

1.228 103

(+34.47%)

analytical
reference

Table 5.2: Numerical results for sandwich structure and loadcase 1

range. The deviation is caused by the discretization of the finite element model
and can be reduced by refining the mesh size. The maximal deflection of the
mass-equivalent reference finite element solution with constant thickness t0 is
significantly higher than the optimized solution.

Numerical Solution
Assuming a full cross-section without core, the bending becomes more important. Thus, a numerical optimization with beam elements has been performed
according to the explanations in Section 5.1.2. In contrast to the sandwich example above, the total reference thickness t0 is set to 20 t pl y . The obtained
solutions are shown in Figure 5.4. Here, the green layers mark the fixed initial layers that are needed to prevent numerical problems of the finite element
model. The solutions are almost identical and the deviation of the maximal de-

0.2

0.4

0.6

0.8

Figure 5.4: Thickness distributions for a single load (numerical)


flection between the two optimization methods is smaller than 1%. Since both
solutions base on finite elements, the deviation caused by the discretization is
very low. The maximal displacement of the mass-equivalent reference solution
is 66.96% higher.

5.2 Minimal Compliance

123

GLM

vmax

29.948 103

vmax

29.703 103

(0.82%)

vmax

50.000 103

(+66.96%)

BEAM FEM
reference

Table 5.3: Numerical results for full cross section and loadcase 1

5.2.2

Loadcase 2: Distributed load

Analytical Solution
In order to compare the finite element designs with a more complex thickness
distribution which is not simply linear, the same calculations are done for another sandwich beam with a uniform distributed load q= FL (see Figure 5.5).
The bending moment caused by the distributed load is
t

q= F
L

E , b, L

Figure 5.5: Sandwich cantilever beam with distributed load

M b (x) =

q
(L x)2
2L

(5.27)

As before, the dummy bending moment M b must be defined which is dependent


on a unit force at the point where the deflection v is measured. Again, the
deflection at the left right side (x = L) is taken wherefore

M b (x) = (L x)

(5.28)

However, the GLM designs with respect to the potential of external forces
which is here
W=

F
L

L
0

v(x)dx

(5.29)

124

Verification

This difference results in a slight error which is discussed in more detail in


Section 5.3. The optimality condition (5.11) forms to
M b (x) M b (x)
(EI z (x))2

q
2L

(L x)3

(E I z (x))2

=C

(5.30)

and the area moment of inertia is




I z (x) =

1
2

q
2L

1
(L x) 2
= h2 bt(x)
E
2

(5.31)

In contrast to the loadcase with a single force, the optimal thickness distribution is not linear anymore.


t(x) =

q
2L

1
2

3
 
(L x) 2 1 2 2
E
2b h

(5.32)

The constant parameter C can be calculated analogously to steps (5.20) to


(5.22).

2
8
q L5
C=
(5.33)
5 E V0 2L h4
The combination of equations (5.32) and (5.33) delivers the optimal thickness
distribution for an uniformly loaded sandwich beam.
3

t(x) =

5(L x) 2 V0
5

4b L 2

5 
x  32
t0 1
2
L

(5.34)

The thickness distribution arising from the GLM and the analytical solution
are shown in Figure 5.6. Again, they match quite well even if the objective
functions are not exactly the same. The deflection along the beam axis is








4Lq 2x2 1 Lx + Lx 5 4 1 Lx + 2L2 1 + 1 Lx
(5.35)
v(x) =
75bEh2 t0
and the maximal deflection on the right side of the beam is
vmax = v(L) =

4L3 q
25bEh2 t0

(5.36)

The deviation between analytical and GLM solution is in the range of the first
example.

5.2 Minimal Compliance

0.2

125

0.4

0.6

0.8

Figure 5.6: Thickness distributions for a uniform distributed load (analytical)


GLM

vmax

0.294 103

vmax

0.320 103

(+9.03%)

vmax

0.464 103

(+57.96%)

analytical
reference

Table 5.4: Numerical results for sandwich structure and loadcase 2

Numerical Solution
The optimal thickness distribution for a full cross section beam with distributed load seems to be linear (see Figure 5.7). Also this solution is found
quite exactly using the GLM. The numerical deviations are in the same range

0.2

0.4

0.6

0.8

Figure 5.7: Thickness distributions for a uniform distributed load (numerical)


as before and listed in Table 5.5. The stiffness increase due to the variable
thickness distribution is relatively high. The bending moment induced by
the distributed load on the right side of the beam is small (and goes to zero

126

Verification

for x = L) wherefore material can be reduced and transferred to the more highly
loaded right side.
GLM

vmax

8.635 103

vmax

9.299 103

(+7.68%)

vmax

18.751 103

(+117.14%)

BEAM FEM
reference

Table 5.5: Numerical results for full cross section and loadcase 2

In summary, the obtained solutions with the GLM are very similar to the
analytically derived optimal solutions. The deviation between the methods is
in an acceptable range and arises primarily from the discretization of the finite
element model. Compared to the numerical solutions with bending load, the
deviation is small. In all examples, the maximal deflection can be reduced
significantly by taking advantage of a variable thickness distribution.

5.3 Minimal Deflection


As mentioned before, a slight error arises in the calculation of the second loadcase from the purposely wrongly chosen objective function. While the GLM
minimizes the potential of external forces W, the analytical solution provides
a minimal solution for the deflection at the left end of the bar. However, the
GLM can be modified to consider the deflection on the right side v(L) as objective as well. To do so, the sensitivities which express the influence of a
thickness change to the deflection v(L) are determined. The resulting distribution is shown in Figure 5.8. The shape differs only slightly from the thickness
distribution obtained with the potential of external forces (Figure 5.6). Also
max
= 0.0293 103 is not significantly smaller.
the maximal deflection of vGLM
The big advantage of using the potential of external forces for the optimization is the fact that no expensive displacement sensitivities are required. If a
structure has to be designed for maximal stiffness, using the potential of external forces is appropriate. Using the deflection as objective function may be
useful for some selected problems, e.g. if the relative displacement between
two points should be minimal.

5.4 Eigenfrequency

127

0.2

0.4

0.6

0.8

Figure 5.8: Thickness distributions for a uniform distributed load (Minimal


Deflection)

5.4 Eigenfrequency
The GLM for the eigenfrequecy optimization is verified by using a bar with a
fixed end and a lumped mass M on the free end according to Figure 5.9. All
b
t
E , b, L ,

Figure 5.9: Bar with lumped mass


lateral degrees of freedom are fixed so that only a motion along the axis of the
bar is allowed. In contrast to the analytical solutions for static loadcases, a full
rectangular bar with variable thickness t(x) and constant width b is considered
which consequently leads to a variable cross section area A(x). The equation of
motion for a bar with variable cross section area A(x) is



u(x, t)
2 u(x, t)
E A(x)
(5.37)
= A(x)
x
x
t2
Assuming a harmonic vibration
sin( t)
u(x, t) = u(x)

and using the wave propagation speed c which is defined as


)
E
c=

(5.38)

(5.39)

128

Verification

gives the following ordinary differential equation:





u(x)
2

A(x)
=0
+ 2 A(x)u(x)
x
x
c

(5.40)

Assuming an unknown variable cross section A(x), the equation cannot be


solved directly.
Analytical Solution
For the design for maximum fundamental frequency, Prager and Taylor [134]
present an optimality criterion which is based on the same assumptions as one
presented in Section 5.2. For this case, the specific stiffness s(x) is the product
of the Youngs modulus E and the the cross section area A(x).
s(x)bar = Ebt(x)

(5.41)

Using the Rayleighs principle it is shown that a design reaches maximum frequency if equation (5.42) is fulfilled.
b2 2
C2
u (x)2 =
(5.42)
2
2
Herein, e denotes the specific strain energy stored in a structural element of
unit stiffness that has been subjected to the strain . The parameters b and C
are constants and u(x) is the axial displacement. In case of a bar, the strain
energy per unit stiffness e is half the square of the strains.

e (q (x))

1
(5.43)
u(x)2
2
Here, the parameter b is set to the inverse of the wave propagation speed in
the bar c. The optimality condition therefore requires that
e(x) =

u(x)2

u(x)2 = C2
(5.44)
c2
Using the boundary condition u(0) = 0, the differential equation yields
 x 
c
(5.45)
u(x) = C sinh

c
This solution is then substituted in to the basic differential equation (5.40).
#
 x $ 2
 x 

c
A(x) Ccosh
(5.46)
+ 2 A(x) C sinh
=0
x
c

c
c
 x $
#
 x $
 x 
A(x) #
sinh
cosh
(5.47)
+ A(x)
+ A(x) sinh
=0
x
c
c
c
c
c
 x 
 x 
A(x)

cosh
+ 2A(x) sinh
=0
(5.48)
x
c
c
c

5.4 Eigenfrequency

129

Solving the differential equation (5.48) for the unknown area A(x) leads to
A(x) =

const
 
cosh2 cx

(5.49)

An additional boundary condition is formulated with the equilibrium of momentum on the right side of the bar.

= A(L)(L)
M u(L)
2

M u(L) = A(L)Eu (L)

(5.50)
(5.51)

The variable area A(x) design for maximum fundamental frequency is therefore
 
 
cM sinh cL cosh cL
(5.52)
A(x) =
 
E cosh2 cx
and the variable thickness t(x) assuming a constant width b is
 
 
cM sinh cL cosh cL
t(x) =
 
Eb cosh2 cx

(5.53)

Calling for volume constancy, the implicit equation for the angular frequency
can be written as





1 L
c2 M
L
2L
t(x)dx =
sinh
(5.54)
tanh
t 0 () =
L 0
2EbL
c
c
This equation cannot be rearranged to an explicit expression of . However, it
can be used to determine the frequency numerically or graphically.
Two different bars with variable thickness for maximal fundamental frequency have been designed using the GLM. In order to ensure resolvability of
the finite element equation system, two layers are predefined which are assumed to be fixed. The mass is constrained to 0.02+ M which corresponds to
a beam with a constant thickness of 20 t pl y . In the first loadcase, a lumped
mass M of 0.02 has been used which is equal to the bar mass of the final design and therefore contributes half of the total mass. Taking advantage of
relation (5.54), the corresponding angular frequency can be determined. This
can be done either numerically or graphically as it is shown in Figure 5.10.
Entering the determined value of in equation (5.53) yields the optimal thickness distribution for the current loadcase. The analytical and the GLM solution are shown in Figure 5.11. The predefined fixed layers are colored green.

Verification

Reference Thickness t0

130

x 10-3

2
1.5
1
2

2.2

2.4
2.6
Angular Frequency

2.8

3
4

x 10

Figure 5.10: t0 --plot M = 0.02 ( = 27.871 103 )

0.2

0.4

0.6

0.8

Figure 5.11: Thickness distributions for a bar with lumped mass M = 0.02
GLM

29.196 103

27.871 103

(4.54%)

28.520 103

(2.32%)

analytical
reference

Table 5.6: Numerical results for a bar with lumped mass M = 0.02

Obviously, the correlation between analytical and GLM solution is high. Due
to the high amount of additional mass on the right side, the thickness relief is
only small. The numerical results are listed in Table 5.6. Again, the solution
obtained with the GLM is slightly stiffer than the analytical solution. However, comparing the optimized solution with a reference finite element design
with constant thickness t0 , the frequency is not increased significantly. The

5.4 Eigenfrequency

131

Reference Thickness t0

relatively large lumped mass M limits the potential for improvement. If the
lumped mass M is much greater then the mass of the bar, the solution with
constant thickness becomes optimal.
In the second loadcase, the lumped mass M is reduced to 0.004 in order for
the self-weight of the bar becoming more important. Theoretically, the lumped
mass could even be removed. However, the optimal design would consequently
concentrate the mass to the very left end of the bar and the thickness on the
right side would be reduced to zero. Such a solution is infeasible for the FEM
since the zero thickness causes a singularity. The angular frequency is again
determined using the graph in Figure 5.12. Also in this case, the correlation
x 10-3
2
1.8
1.6
1.4
1.2
4

4.2

4.4
4.6
Angular Frequency

4.8

5
x 104

Figure 5.12: t0 --plot M = 0.004 ( = 48.841 103 )


between analytical and optimized solution is high (see Figure 5.13). In contrast

0.2

0.4

0.6

0.8

Figure 5.13: Thickness distributions for a bar with lumped mass M = 0.004
to the first example, the thickness relief is apparent. This results also in a
higher frequency compared to the reference finite element solution (see Table
5.7) with constant thickness t0 .

132

Verification

GLM

51.073 103

48.841 103

(4.37%)

43.553 103

(14.72%)

analytical
reference

Table 5.7: Numerical results for a bar with lumped mass M = 0.004
Numerical Solution
Here as well, the example above considers no bending load wherefore a second
loadcase is calculated. The first fundamental frequency of a hinged full-crosssection beam with an eccentric lumped mass is maximized (see Figure 5.14).
The reference thickness t0 is set to 20 t pl y which yields a structural mass
of 0.02. The eccentric mass, which is also 0.02, causes an unsymmetric mode
which results in an unsymmetric thickness distribution. The numerical sob
t
E , b, L ,

Figure 5.14: Hinged beam with eccentric lumped mass


lution with beam elements and the GLM solution are almost identical (see
Figure 5.15). This is confirmed by the numerical results in Table 5.8. The de-

0.2

0.4

0.6

0.8

Figure 5.15: Thickness distribution of the hinged beam


viation of the angular frequencies for the two methods is less than 1%. The
frequency increase due to the variable thickness is significant considering that
the lumped mass contributes 50% of the total mass. The coincidence between

5.5 Buckling Load Factor

133

GLM

156.462

157.014

(+0.35%)

136.094

(13.02%)

BEAM FEM
reference

Table 5.8: Numerical results for beam with eccentric lumped mass M = 0.02
analytical solutions, and numerical beam solutions respectively, and the solutions obtained with the GLM is also high for dynamic problems.

5.5 Buckling Load Factor


Prager and Taylor [134] present a further optimality criterion which postulates
that a sandwich beam with optimal buckling load to mass ratio has a constant
curvature.
v (x) = C

(5.55)

To verify the presented method for the optimal design of structures for buckling
load, a simple hinged sandwich beam loaded with an axial force F is considered
(see Figure 5.16).
t
F

E , b, L

Figure 5.16: Hinged beam with single axial force

Analytical Solution
Solving this differential equation (5.55) and using the boundary conditions for
the hinges (v(0) = 0 and v(L) = 0) yields the deflection
u(x) =

C
x (L x)
2

(5.56)

134

Verification

The bending moment M b is dependent on the critical load P

(5.57)

M b (x) = Pv(x)

and must be equal to


M b (x) = s(x)v (x)

(5.58)

whereas s(x) is the product of E and I z (x). The combination of these equations
yields

P
1
(5.59)
x (L x) = Eh2 bt(x)
2
2
which can be solved for t(x).
P
(5.60)
x (L x)
t(x) =
Eh2 b
The thickness t(x) is dependent on the critical P. Calling for volume constancy
L
(5.61)
V 0 = bLt0 = b
t(x)dx
s(x) =

the critical load P can be determined.


6bEh2 t0
(5.62)
L2
The buckling load factor cr expresses the potential load increase until buckling occurs and is defined as

P=

P
(5.63)
F
The thickness distribution of the considered sandwich beam is illustrated in
Figure 5.17. Also for the buckling case, the agreement between the solution
obtained with the GLM and the analytical solution is good. The numerical results are shown in Table 5.9. Obviously, the buckling load factors for analytical
and GLM are almost equal. The buckling load factor of the reference solution
is significantly lower.
=

Numerical Solution
For the verification of a buckling loadcase with bending influence, a full cross
section beam which is clamped on the left side and hinged on the right side
is loaded with an axial force. The chosen boundary conditions lead to an unsymmetric bending moment whereas the thickness distribution will be unsymmetric as well. The obtained thickness distributions are shown in Figure 5.18.
Also in this case, the two solutions coincide fairly well. The deviation of the
buckling load factors is only 1.74%.

5.5 Buckling Load Factor

135

0.2

0.4

0.6

0.8

Figure 5.17: Buckling optimal thickness distribution for a hinged beam


GLM

3.075 103

3.000 103

(2.44%)

2.566 103

(16.55%)

analytical
reference

Table 5.9: Numerical results for hinged beam

b
F

E , b, L

Figure 5.18: Clamped-hinged beam with single axial force


GLM

187.834

191.107

(+1.74%)

134.629

(28.33%)

BEAM FEM
reference

Table 5.10: Numerical results for clamped-hinged beam with single axial force

136

Verification

0.2

0.4

0.6

0.8

Figure 5.19: Thickness distribution for clamped-hinged beam with single axial
force

5.6 Strength
It is difficult to prove analytically that a structure is optimal with respect to
strength or not. Having isotropic materials, the optimality criterion of fully
stressed design is usually used [113, 142, 52]. It postulates that a structure
has an optimal material distribution in terms of strength if the stresses are
distributed homogeneous through the material. If the stresses reach the limit
at every point of the structure, material can be reduced nowhere without causing failure. This criterion appears plausible but there exists no mathematical
proof for general validity. Analytical solutions with a fully stressed design are
those of the sandwich cantilever beam for minimal compliance in Section 5.2.
There, the stresses in the facesheets are constant. However, such solutions
are for isotropic materials. Here, only strength criteria for composites, namely
Tsai-Hill and Tsai-Wu have been implemented. The method thus cannot be
verified with respect to strength using an analytical solution as done above.

5.7 Discussion
The verification examples above demonstrate that the GLM is able to approximately find solutions which are proved to be optimal. The analytically derived
solutions and the solutions obtained with the GLM are almost identical except for the difference caused by the discretization errors. The GLM has been
shown to reproduce the optimum solutions which are analytically described for
sample structures made of isotropic materials. Analytical solutions for problems including the parameters of anisotropic layered materials, with which the
GLM could be verified, are not available.

Chapter 6

Laminate Applications
This chapter demonstrates the presented method for a selection of applications
with laminates. The first example demonstrates an eigenfrequency optimization of a laminated panel with large cutouts. Different objective functions such
as maximization of the eigenfrequencies but also more complex schemes for
finding search directions for the objective of splitting eigenfrequencies are applied. Nodal displacement sensitivities are used to design a compliant grabber.
A third example explores the method for buckling applications for two plate
geometries. Consequently, the method is applied on a strength problem of a
plate with a centered hole. Stress concentrations are reduced with local reinforcements which results in a significantly increased strength. The chapter
is terminated with the optimization of an automotive component regarding to
stiffness and mass.

6.1 Vibrating Panel (Eigenfrequency)


The potential of the proposed process is illustrated on an eigenfrequency optimization for a carbon fiber reinforced panel of 400 times 300 millimeters shown
in Figure 6.1. The stiffness of the plate is reduced due to three big rectangular
cut-outs. Additionally, a mass of 10 g is attached to the center. Four spring
elements are attached at the corners of the panel and fixed on the other end.
They simulate the rubber cords which have later been used for real measurements. However, the model is not statically determined wherefore the rigid
body modes close to a frequency of zero occur. They are filtered in all calculations. The panel consists of a symmetric, quasi-isotropic laminate with
layup (0/45/-45/90)S made of unidirectional carbon fiber reinforced epoxy (see

138

Laminate Applications

400
60

120

30

150
300

100

50

30

80

y
x

50

=0

60

200

Figure 6.1: Design domain of a panel with cut-outs


Table A.2). Each ply has a thickness of 0.25 mm which results in a total thickness of 2 mm and a total mass of 0.287 kg. The finite element model consists of
3560 CQUAD4 elements which results in 22788 degrees of freedom. The lowest
6 modes and the corresponding eigenfrequencies of the basic panel are shown
in Figure 6.2. In the following Sections, the panel is reinforced locally using the
proposed method considering different objective functions. Four ghost layers
with orientation angles (0/45/-45/90) are attached on both sides. The laminate
is symmetric wherefore only the stack-up of one symmetric half needs to be
included in the parameter set. The reinforcement layers are generated until a
mass of 0.3834 kg is reached. This is exactly the mass of the panel with same
dimensions and layup but without cut-outs. The step-size is controlled with a
2%-threshold and a smooth factor of 2 is chosen for all examples.

6.1.1 Maximization of the Lowest Eigenfrequency


A potential objective of a panel design with local reinforcements could be the
maximization of the lowest eigenfrequency. The lowest eigenfrequency is sometimes used as a measurement for the stiffness of a structure, e.g. in space applications. The fundamental natural frequencies of a satellite structure must
exceed a given threshold in order to prevent dynamic coupling between satel-

6.1 Vibrating Panel (Eigenfrequency)

139

(a) f1 = 47.45 Hz

(b) f2 = 76.63 Hz

(c) f3 = 82.00 Hz

(d) f4 = 119.61 Hz

(e) f5 = 137.03Hz

(f) f6 = 188.14 Hz

Figure 6.2: Lowest six eigenmodes of the basic panel with corresponding freqiencies
lite and launcher, while the mass should be as low as possible. To increase
the lowest eigenfrequency, only the sensitivity field of lowest eigenmode using
equation (3.53) is considered. The search direction s thus yields
s=

d 1
dt

(6.1)

A similar example with a slightly different initial configuration has been presented in [145]. The change of the first six frequencies during the patch genera-

140

Laminate Applications

300

Eigenfrequency [Hz]

250

200

150

100

50

0.29

0.3

0.31

0.32

0.33
0.34
Mass [kg]

0.35

0.36

0.37

0.38

Figure 6.3: Frequency characteristic: maximization of the lowest eigenfrequency

tion process is shown in Figure 6.3. Even if the modes 2 to 6 are not considered
here, they increase as a consequence of the stiffening regarding to mode 1. Notice that the first and the second eigenfrequency merge during the process.
The first frequency becomes slightly larger than the second one. Since the algorithm does not take care of the mode shape, the structure is reinforced considering the second mode until it is again higher than the first one. This alternation can be understood as a mode switch. Since the iteration steps are chosen small, it seems that the frequencies are coinciding. However, the two mode
shapes are different. The terminal solution is reached after 457 iterations. The
lowest frequency of the locally reinforced panel is raised to 104.42 Hz which is
120.05% higher than the frequency of the initial layup. Thereby, the mass was
increased by only 34.29% (which is given by the termination criterion defined
above). A complete overview of the results is given in Table 6.1. The local reinDesign
init
opt
opt ( % init )

f1

f2

f3

f4

f5

f6

47.45
104.42
120.05

76.63
104.65
36.58

82.00
153.60
87.32

119.61
206.51
72.65

137.03
213.09
55.51

188.14
260.39
38.40

0.2870
0.3854
34.29

Table 6.1: Result table: maximization of the lowest eigenfrequency

6.1 Vibrating Panel (Eigenfrequency)

141

forcements for the four different orientation angles are shown in Figure 6.4. It
is obvious that such contours could never be found intuitively. There are only
a few areas with slightly scattered reinforcements which have to be mitigated
manually. However, the connecting areas prevail wherefore manufacturing of
the design solution is feasible. The advantage of locally reinforced laminates

(a) = 90

(b) = 45

(c) = 45

(d) = 0

Figure 6.4: Reinforcement contours: maximization of the lowest eigenfrequency


becomes obvious considering Figure 6.5. There, the run has been continued
until all ghost layers have completely turned to real layers. Thus, the terminal solution consists of fully covering layers which is equivalent to a solution
with a doubled basic layup. The mass is almost doubled (not exactly because of
the point mass) and its lowest frequency is 102.07 Hz which is slightly below
the lowest frequency of the above obtained solution. The overshooting of the
frequency emphasizes that locally reinforced solutions have remarkable better

142

Laminate Applications

frequency-to-mass ratios than fully covering designs. For all six fundamental
frequencies, the maximum value is reached before all the potential reinforcement material is used. This is especially the case for the first frequency since
this is the considered objective function. This effect may be smaller for higher
modes since they have many local deformations. However, higher modes are
often not important for structural design. Another interesting phenomenon
can be observed in Figure 6.5. The same first frequency occurring at the fully
covering design can be reached by using only 32.75% of the potential reinforcement material which implies that much of the material of the fully covering
solution is useless.
450
f

400

Eigenfrequency [Hz]

350
f

300
250

max
6

max
max f5
4

max
3

200
f

150
f
100

max
f
1

eq

max
2

50
0

0.3

0.35

0.4
0.45
Mass [kg]

0.5

0.55

Figure 6.5: Frequency characteristic: maximization of the lowest eigenfrequency; complete run

6.1.2 Maximization of -Weighted Eigenfrequencies


Instead of performing a maximization of only the lowest eigenfrequency, a
weighted sum of the sensitivities of the lowest frequencies is considered in
the second optimization. The weighting factors could theoretically be chosen
arbitrarily. A constant or linear distribution is possible. Here, the sensitivity values are weighted with the reciprocal of the square of the corresponding
eigenvalue. The search direction is
s=

n 1 d

n
2
dt
i =1 n

(6.2)

6.1 Vibrating Panel (Eigenfrequency)

143

The purpose of this approach is to weight the lower frequencies higher so that
they are raised more. The effect of this weighting is shown in Figure 6.6. In
300

Eigenfrequency [Hz]

250

200

150

100

50

0.29

0.3

0.31

0.32

0.33
0.34
Mass [kg]

0.35

0.36

0.37

0.38

Figure 6.6: Frequency characteristic: maximization of -weighted eigenfrequencies


contrast to the optimization considering only the lowest mode, it can be avoided
that the first and the second frequency fall together. The terminal solution is
reached after 280 iterations. The results are summarized in Table 6.2. The inDesign
init
opt
opt ( % init )

f1

f2

f3

f4

f5

f6

47.45
97.80
106.11

76.63
127.13
65.90

82.00
174.95
113.36

119.61
222.20
85.77

137.03
231.54
68.98

188.14
298.13
58.46

0.2870
0.3854
34.29

Table 6.2: Result table: maximization of -weighted eigenfrequencies


crease of the lowest eigenfrequency is slightly lower compared to the optimization before. However, the upper frequencies are all higher than before. The
reinforcement contours are displayed in Figure 6.7. Comparing them to the solution above the shape is rather different. The reinforcements are concentrated
increased at the center of the structure which is caused by the consideration of
all modes.

144

Laminate Applications

(a) = 90

(b) = 45

(c) = 45

(d) = 0

Figure 6.7: Reinforcement contours: maximization of -weighted eigenfrequencies

6.1.3 Eigenfrequency Splitting


The weighting factors can be chosen arbitrarily so that the potential for structural design is very wide. This optimization aims to split the second and the
third frequency. Thus, the weighting factor for the sensitivities of the second
mode is -1 and +1 for the third one. The search direction is
s=

d 2 d 3
+
dt
dt

(6.3)

The frequency characteristics during the optimization are plotted in Figure 6.8.
It is obvious, that the second and the third frequency drift apart from each
other. Compared to the initial solution, for which the eigenfrequencies are
distributed regularly, the reinforced structure has a resonance-free band between 79.15 Hz and 208.80 Hz. This opens new possibilities for the design of

6.1 Vibrating Panel (Eigenfrequency)

145

300

Eigenfrequency [Hz]

250

200

150

f = 79.15 ... 208.80 Hz

100

50

0.29

0.3

0.31

0.32

0.33
0.34
Mass [kg]

0.35

0.36

0.37

0.38

Figure 6.8: Frequency characteristic: frequency splitting [f2 | f3 ]

the panel. It may be interesting for applications where a source excites the
structure at a specific frequency. Usually, the structure is designed to push all
eigenfrequencies above the excitation frequencies. However, this may generally require more stiffening which increases the weight. If there is no reason
allowing the eigenfrequencies to be below the excitation frequency, this may be
an alternative. Recalling the objective of the optimization, it must be noticed
that the second eigenfrequency is not dropping (it is even rising by 3.29%).
This is caused by the fact that the design process here only allows to add material. While the reinforcements are generated, the second frequency drops since
the reinforcement material contributes to the mass but not to the stiffness of
mode 2. Nevertheless, with the growing of the reinforcements, the eigenfrequency starts to rise again. There is also the possibility to enable the process
to reduce material of the basic plate as it is shown later. A complete overview
of the results is given in Table 6.3 and the reinforcement contours are shown
in Figure 6.9. The resulting reinforcement patterns are complex and cannot
Design
init
opt
opt ( % init )

f1

f2

f3

f4

f5

f6

47.45
71.44
50.56

76.63
79.15
3.29

82.00
208.80
154.63

119.61
211.27
76.63

137.03
226.12
65.02

188.14
267.52
42.19

0.2870
0.3854
34.29

Table 6.3: Result table: frequency splitting [f2 | f3 ]

146

Laminate Applications

(a) = 90

(b) = 45

(c) = 45

(d) = 0

Figure 6.9: Reinforcement contours: frequency splitting [f2 | f3 ]


be found by intuition which emphasizes the value of the method for practical
design. Taking a look at the modes 2 and 3 of the initial laminate configuration
shown in Figure 6.2, it can be noticed that their main deformation is roughly
perpendicular. The stiffening of unidirectional reinforcements in one direction
is more than one order of magnitude higher than in perpendicular direction
which is advantageous for splitting these modes. Considering the reinforcements, it can be noticed that the 90 -layer has developed over the full width of
the panel. This stiffens mainly the third mode while the stiffness contribution
to the second mode is rare. Almost no reinforcement can be observed on a vertical line through the left central cut-out which guarantees the compliance of the
second mode and keeps it low, respectively. However, the same simulation has
been carried out splitting the eigenfrequencies 3 and 4 (shown in Figure 6.10).
Also there, a resonance-free band can be identified even if it is not as wide as
before. Thus, any two eigenfrequencies can basically be split but the success

6.1 Vibrating Panel (Eigenfrequency)

147

might be different.
300

Eigenfrequency [Hz]

250

200
f = 124.79 ... 239.45 Hz
150

100

50

0.29

0.3

0.31

0.32

0.33
0.34
Mass [kg]

0.35

0.36

0.37

0.38

Figure 6.10: Frequency characteristic: frequency splitting [f3 | f4 ]


Using the splitting method above, a resonance-free band can be achieved easily
but cannot be controlled in which range the band occurs. A better way to control this is to use the weighting factors of equation (4.7) (see Figure 4.5). The
search direction becomes
s=

n


1
d n
i=1 i thres dt

(6.4)

Eigenfrequencies that are below the threshold f thres are minimized and frequencies above are maximized. The closer a frequency comes to f thres the
higher is the weighting factor. The reinforcement generation process has been
carried out setting the threshold frequency to f thres =80 Hz. Consequently, the
threshold is between the second and the third mode. In contrast to the optimizations above, the process is allowed to reduce material of the basic plate. It
must be taken care that no regions occur where all layers are removed. This
would of course result in numerical problems in the finite element run. To
avoid that, the 90 -layer of the basic laminate is locked which means that it
cannot be reduced. It is the most important layer for the third mode so that
the result is falsified minimal. Since material can be reduced as well, it may
be impossible to reach a given weight. The termination criterion has been set
manually to 260 iterations. Figure 6.11 shows the characteristics of the eigenfrequencies during the process. In contrast to the examples above, the mass

148

Laminate Applications

240
220

Eigenfrequency [Hz]

200
180
160
140
120
100
80
60
40

0.286

0.288

0.29

0.292
Mass [kg]

0.294

0.296

0.298

Figure 6.11: Frequency characteristic: frequency splitting with threshold


[80 Hz]
does not increase monotonously since it can also be reduced. It is obvious that
the second and the third eigenfrequencies drift apart from the threshold of 80
Hz which is indicated by the black line. In contrast to the runs above, the
second eigenfrequency is dropping significantly.
Design
init
opt
opt ( % init )

f1

f2

f3

f4

f5

f6

47.45
42.06
-11.37

76.63
51.43
-32.88

82.00
131.08
59.85

119.61
175.20
46.47

137.03
193.45
41.18

188.14
224.715
19.44

0.2870
0.2957
34.29

Table 6.4: Result table: frequency splitting with threshold [80 Hz]
The presented examples demonstrate the great potential of locally reinforced laminates considering vibration problems. By choosing an appropriate
search direction, the frequency characteristics of laminated structures can be
designed specifically. The reinforcement contours are complex and cannot be
found intuitively which emphasizes the value of the presented method for practical design.

6.2 Grabber (Nodal Displacement)

149

6.2 Grabber (Nodal Displacement)


A potential application using the sensitivities of nodal displacement has already been introduced in Section 3.4. There, a structure of a soft isotropic
basic material is predefined (see Figure 3.8(a)). The displacement field, which
results from the applied forces on the left hand sides, is shown in Figure 3.8(b)
The two reference points on the right hand side drift apart from each other.
Four ghost layers of a unidirectional reinforced material with orientation angles (0/45/-45/90) are added in order to reinforce the structure in a way that
the reference points are getting closer to each other. The search direction s is
the derivative of the relative displacement between the two reference points.
s=

with


 

re f
re f
re f
re f 2
= y1 + v1 y2 + v2

(6.5)

Consequently, the vertical distance between the two points is minimized. Only
36 iterations are needed until the best solution in terms of minimal distance
between the reference points is obtained. The solution for each reinforcement
layer and for the thickness sum of all layers is shown in Figure 6.12. It re-

45

-45

90
(a) Ply-by-ply

(b) Total thickness plot

Figure 6.12: Local reinforcements of the grabber example


minds of a so-called compliant mechanism which provide their functionality
without taking advantage of any hinges. The displacement field for the reinforced structure is shown in Figure 6.13. The colormap only considers displacement in y-direction. Comparing this solution with the displacement field before

150

Laminate Applications

(see Figure 3.8(b)), it becomes obvious that the reference points on the right
side are now getting closer to each other. Table 6.5 summarizes the relative
y-displacement per unit force between the two reference points. It emphasizes
that the reference points approach each other for the reinforced configuration.
While a homogeneous material cause the reference points to diverge, the specific distribution of local anisotropies reverses the relative displacement. The
Design
basic
reinforced

re f

u y /F [mm/kN]
0.932
-0.106

Table 6.5: Result table: grabber


design seems to be inappropriate from the manufacturing point of view. However, the example demonstrates that the GLM is able to find solutions with the
introduced criterion of nodal displacement. Considering the thickness distributions it becomes obvious that such a design can hardly be found intuitively.
Moreover, the mechanisms that cause these points to approach each other are
not easy to comprehend. The example indicates great potential for using the

Figure 6.13: y-displacements of the locally reinforced grabber


GLM in combination with the criterion for nodal displacement. Alternatively,
the structure could be designed so that the relative displacement between the
reference points is minimal taking the objective function (3.73) which neglects
the nodal coordinates. Due to the symmetrical loading, the displacements of
the reference points become invariant of the applied load. The minimal specific, relative y-displacement of the reference points obtained in an interim
solution of the optimization above is 3.7 m/kN.

6.3 Buckling Plate (Buckling)

151

6.3 Buckling Plate (Buckling)


The GLM considering buckling loads is demonstrated on a simple plate with a
dimension of 300 mm times 200 mm. The basic laminate of the plate is a simple quasi-isotropic with layup (0/45/-45/90)S . It is simply supported at all edges
and it is loaded with pressure in the long-side direction. The forces are chosen
so that the buckling load factor cr is exactly one which means that buckling
occurs. The corresponding buckling mode is shown in Figure 6.14. Goal of the

200

300

Figure 6.14: Geometry, load and first buckling mode of a rectangular plate
optimization is to reinforce the structure locally in order to double the buckling
stability ( cr = 2). The allowed orientations of the reinforcement layers are 0,
60 and -60 degrees. Such a layup is known as the quasi-isotropic layup with
minimal number of layers needed. It assures that the structure can be stiffened
in every direction. All reinforcement layers are attached on both sides which
provides a solution with symmetric laminate. For each iteration loop, the element layers within the highest 10% are adapted and a smooth factor of 2 is
applied. The run is terminated after 42 iterations. According to Section 3.5,
the sensitivities are evaluated with the stiffness matrix only. The characteris-

Buckling Load Factor []

2.5

1.5

0.5
0.185

0.19

0.195
0.2
Mass [kg]

0.205

0.21

Figure 6.15: Characteristics of the first and second buckling load factor
tics of the first and the second buckling load factor during the patch generation

152

Laminate Applications

process are shown in Figure 6.15. All higher buckling load factors are not of importance since they are higher than the targeted value of 2. Moreover, higher
modes cannot even occur since the structure usually fails already when the
first mode is reached. Obviously, the first and second buckling load factors converge during the optimization. As search direction s, only the sensitivities of
the lowest mode are considered wherefore the structure is reinforced to raise
the load factor of the first mode. However, at a certain point a mode switch
occurs which means that the second mode drops below the first one. Consequently, the search direction considers the second one. More mode switches
occur until the optimization is terminated. The contours of the reinforcements
for the three orientation angles are shown in Figure 6.16. Additionally, a plot of

(a) = 60

(b) = 60

(c) = 0

(d) Total Thickness

Figure 6.16: Reinforcement contours: maximization of the buckling load


the total thickness is added. The contours have been modified slightly by hand
in order to make the layers connected. However, the results do not change significantly. The 60 -layers mainly reinforce the first mode which is shown in
Figure 6.14. The 0 -reinforcements mainly increase the second buckling mode

6.3 Buckling Plate (Buckling)

153

Figure 6.17: Geometry, load and first buckling mode of a rectangular plate with
centered cutout

which has two buckles. In order to verify if the chosen orientation angles are
appropriate to find a good solution, a second optimization has been performed
where the reinforcement layers are allowed to have more different orientation
angles, namely (0/30/45/60/90). The results are summarized in Table 6.6.
Even if more orientation angles are allowed, no significant improvement of the
Design

m [kg]

basic
(0/60)
(0/30/45/60/90)

0.187
0.208
0.207

1 [-]

(+11.49%)
(+10.71%)

1.00
2.05
2.06

(+104.96%)
(+105.78%)

Table 6.6: Result table: buckling plate


solution can be realized. The required additional mass to double the buckling
stability is only slightly lower than for the solution with only three orientation
angles. From the manufacturing point of view, the design with three reinforcement orientations is much easier and cheaper to realize.
The buckling stability is today usually increased by stiffening the structures
with stringers. It cannot be answered here definitively whether the presented
solution is lighter than a stringer-stiffened design. If the stringers are attached
with rivets, it might be hard to use less than approximately 10 % additional
mass. If using stringers, they have to range from one supported end to another
in order to take advantage of the stiffening effect. Considering structures with
cutouts, as for example shown in Figure 6.17, the stringers have to be moved
out of the center in order not to end at the cutout edge. More than one stringer
is probably needed which raises the mass additionally. The GLM however automatically takes care of the cutouts when generating the reinforcement layers.
The plate with cutouts above has been reinforced analogously to the example
above. Also here, the goal is to double the buckling stability. The parameters have been chosen equally to the example above. The reinforcement layers
and the thickness plot are shown in Figure 6.18. As done before, a second

154

Laminate Applications

(a) = 60

(b) = 60

(c) = 0

(d) Total Thickness

Figure 6.18: Reinforcement contours: maximization of the buckling load with


cutout
optimization run has been performed with more potential reinforcement orientation angles. The results are summarized in Table 6.7. In contrast to the
example above, the mass of the design with the higher number of orientation
angles is much lower. However, the solution with (0/60) only is again easier to
realize. A stringer stiffened solution would probably not be lighter than it. It
Design
basic
(0/60)
(0/30/45/60/90)

m [kg]
0.171
0.187
0.183

1 [-]

(+8.87%)
(+6.81%)

1.00
2.02
2.01

(+101.55%)
(+101.30%)

Table 6.7: Result table: buckling plate with cutouts


is difficult to estimate if the presented designs are competitive with a stringer
solution from the financial point of view.

6.4 Notched Plate (Strength)

155

6.4 Notched Plate (Strength)


The Pseudo Strength Function (PSF) which has been introduced in Section 3.6
and the application of its sensitivities are illustrated on a strength optimization of a notched plate under uniaxial tensile load. The design arising from the
optimization has later been produced and tested in order to verify the strength
increase gained with the local reinforcements. The experimental results are
shown in Section 8.2. The considered plate sketched in Figure 6.19 has a
200

=0

20

50

y
x

Figure 6.19: Design domain of the notched plate


dimension of 200 mm times 50 mm and is made of unidirectional reinforced
carbon/epoxy prepregs (see Table A.2) with a symmetric layup of (0/45/ 45)S .
The thickness of a single ply is 0.25 mm which results in a total thickness
of 1.5 mm. The hole has a diameter of 20 mm and is located at the center of
the plate. First Ply Failure (FPF) occurs at a load of P FPF = 11136 N when
predicted with the Tsai-Hill criterion. Even if there exist more sophisticated
and exact failure criteria for the prediction of composite failure the Tsai-Hill
criterion is assumed to be adequate for these investigations. As mentioned
in Section 3.6.2, the Tsai-Hill criterion is predestined for the calculation of
sensitivities due to its simplicity. Moreover, the objective of this optimization
is to double the strength of the initial configuration which makes the absolute
load for which damage occurs not important.
For the optimization, the load is set to twice the FPF-load
(POPT = 2P FPF = 22272 N). The optimization is performed until all failure indices drop below the critical value of 1. The goal of the optimization is
to double the strength of the notched plate by using local reinforcements. The
orientation angles of the potential reinforcements are predefined by adding
three ghost layers with orientation angles of (0/45/-45) on both sides of the
laminate. Assuming that all ghost layers turn to real layers, the optimized
plate would have the double layup of the initial plate and thus be theoretically
twice as strong. Therefore, a failure load of 22272 N will be reached anyway
during the patch generation process. However, the goal is to reach it with a

156

Laminate Applications

significantly lower mass. In order to reduce the computational costs, only a

Figure 6.20: Mesh of the quarter model with equivalent boundary conditions
quarter model of the notched plate is considered which is shown in Figure 6.20.
Boundary conditions have to be adjusted to model the double symmetry. The
nodal displacements in x-direction of the left cutting edge as well as in
y-direction of the lower cutting edge are locked. Additionally, the out of plane
rotational degrees of the cutting edge nodes are locked. Also the applied total
tensile load has to be divided by a factor 2. To prevent misunderstandings, any
mass and load values here are given with respect to the full model. Figure 6.21

Failure Index (Tsai-Hill) [-]

-400
2

-600

1
-800

3
Mass Increase [%]

-1000
6

Figure 6.21: Characteristics of the optimization process

Pseudo Strength Function Value [-]

-200

6.4 Notched Plate (Strength)

157

shows the characteristics of the optimization process. The x-axis quantifies


the mass percentage in which 0% identifies the initial solution and 100% the
fully developed solution with double layup. The maximal failure index of the
current solution is displayed with the thicker red solid line and quantified on
the left hand y-axis. The initial failure index has a value of 2 since two times
the FPF-load is applied. At the beginning of the optimization, the maximal
failure index starts to increase significantly up to a value of 2.32. This can
be explained with the fact that at the beginning, only a few elements are
covered with a 0 -layer in the regions with highest stresses. The resulting
stiffness discontinuity causes additional stress concentrations according to the
investigations in Section 4.5.3. However, the optimization algorithm moves
forward, even if some failure indices increase. It considers the sensitivities of
the PSF which are dependent on the failure indices at all locations. The PSF
is indicated with the green dash-dotted line and quantified on the right hand
y-axis. It increases continuously and asymptotically towards the theoretically
maximal value of zero which will however never be reached for most of the
structures. Also the average failure index, which is indicated by the thin
line, drops continuously. After the initial increase, the maximal failure
index drops asymptotically towards a failure index of one. This is plausible
since the fully developed solution must have a maximal failure index of 1.
However, the curve undershoots the critical value already at a significantly
lower mass. The optimized solution has a strength equivalent to the fully
developed solution, but it utilizes only 6.00% of the potential reinforcement
material. In absolute terms, the optimized solution takes only 53.00% of the
fully developed solution. This is a remarkable mass saving and emphasizes
the potential of the proposed method. The optimization terminates after 548
iteration loops. The number of iteration loops could be reduced by thickening
more elements per step. To get an accurate design for the experiments, the
step size has been kept low here. Elements whose sensitivities are within 2%
of the maximal sensitivity are adapted. Probably, even lighter solutions could
be found if the basic layup would be treated as variable. The results of the
generated designs are summarized in Table 6.8. Figure 6.22 shows the shape
Design
init
opt
full

PFPF [N]

m [g]

mminit
mfull minit [%]

11136
22272
22272

22.61
23.96
45.21

0
6.00
100

m
mfull [%]

50
53.00
100

Table 6.8: Result table: notched plate


of the reinforcement layers of the terminal solution. To make details visible,

158

Laminate Applications

(a) = 45

(b) = 45

(c) = 0

Figure 6.22: Reinforcement layers of the notched plate

only parts where reinforcement layers have developed are shown. Obviously,
the major strength increase arises from the 0 -layer. The two 45 -layer
have only developed close to the hole edge. It might seem that they are not
really useful for the strength increase. However, it must be considered that
the plate is optimized considering FPF which is in this case likely caused by
matrix cracking. The 45 -reinforcements make a significant contribution
to the load carrying capacity of the plate. Due to the chosen element-based

6.4 Notched Plate (Strength)

159

parameterization scheme, the reinforcement layers may be scattered. This is


the case in some areas of the 0 - and the 45 -layer. However, the connecting
parts prevail so that the solutions can be manufactured without changing the
design significantly.
Figure 6.23 schematically illustrates the conversion of the element-based
representation of the 0 -doubler to a design that is producible. The rough shape

Figure 6.23: Doubler conversion to a production-feasible solution


of the doubler leads to a continuous contour that can be cut manually by a textile cutting machine. Since the design process considers only a quarter model,
the doubler has to be mirrored. The validation of this example is presented
in Section 8.2. The manufacturing process of the specimens is presented in
Chapter 7.
2%

1%

0%

(a) basic configuration


2%

1%

0%

(b) reinforced configuration

Figure 6.24: Von Mises strain distribution

160

Laminate Applications

Figure 6.24 opposes the Von Mises strain field of the basic configuration
with the strain field of the reinforced configuration in order to give an impression of the strain release due to the reinforcements. While the maximum
occurring Von Mises strains in the basic plate are 1.90%, they are reduced to
1.63% with the reinforcements, which seems not to be so much. However, the
reinforcements do not only change the stress magnitude but also the principal directions wherefore the strongly anisotropic material is loaded more appropriately. Considering the Tsai-Hill Failure Indices of the critical layers in
Figure 6.25, which is the +45 -layer for the basic and the 45 -layer for the reinforced configuration, it can be recognized that the maximum value is reduced
by a factor of 2.
2

(a) basic configuration (= +45 )


2

(b) reinforced configuration (= 45 )

Figure 6.25: Tsai-Hill Failure Indices of the critical layer

Chapter 7

Manufacturing-Related
Aspects
The locally reinforced laminate designs obtained with the proposed GLM are
of relatively high complexity. The manufacturing of such parts is more expensive compared to conventional laminates. However, the outstanding structural
response may mitigate the drawbacks originating in the more complex and
more expensive manufacturing process since the overall lifetime cycle becomes
cheaper.
There exist a couple of composite manufacturing processes that are able
to realize the reinforced solutions. One class of these are the Liquid Composite Molding (LCM) processes to which the well-known Resin Transfer
Molding (RTM) process belongs to. They are highly capable for mass production wherefore they are widely spread in the industry. The manufacturing of
molds for composite designs obtained with the GLM and the additional equipment are expensive. The mold shapes have to vary according to the thickness
variation of the composite parts. Such a process is only feasible if many parts
have to be manufactured.
Within this thesis, only a few specimens are manufactured in order to validate the obtained simulation results. An appropriate manufacturing process
for low-piece-number composites with good laminate quality and dimensional
accuracy is the PREPREG-technique. PREPREGs are unidirectional or woven
fibers which are pre-impregnated with the matrix system which is typically an
epoxy resin. They are stacked in an uncured state and usually cured in an autoclave process. The structure is enclosed in a vacuum bag in order to remove
voids and improve the laminate quality, namely the fiber volume content. Ad-

162

Manufacturing-Related Aspects

ditional external pressure is applied in order to further increase the laminate


quality. Dependent on the matrix system, the laminate thickness, the temperature and the curing degree needed, the process takes between a few minutes
and a few hours. Basically, also the PREPREG-technique requires a mold which
defines the shape of the final part. In case of plane and constant-thickness
laminates, simple plates can be used. To be able to produce the locally varying thickness designs without the need of complex molds, a sequential curing
process is suggested.

7.1 Sequential Curing Process


Many composite parts consist of a symmetric layup due to various reasons.
Symmetric laminates have no coupling between bending and membrane deformation. Thus, the calculation is simplified which reduces the analysis effort.
Moreover, membrane-bending coupling is unwanted in many cases. The structural concepts are often derived from designs which have been made of isotropic
materials. Since these materials have no membrane-bending relation, it is unlikely seen for composites as well. Also the here considered specimens have a
symmetric layup.
Figure 7.1(a) schematically illustrates a locally reinforced laminate with
symmetric layup where the black lines represent the single plies. Such a complex thickness distribution can be manufactured using a mold as indicated by
the surrounding parts. To avoid an utilization of expensive molds, a plane plate
can be used by simultaneously taking advantage of a vacuum bag. However,
the symmetry of the laminate is lost which is sketched in Figure 7.1(b). Thus,
a simple vacuum bag technique is infeasible for manufacturing the locally reinforced specimens.

(a) Mold

(b) Vacuum bag

Figure 7.1: Potential PREPREG manufacturing techniques


The here used manufacturing process, also called sequential curing process, takes a different approach. All the considered specimens have in common
that they contain a basic laminate of a constant thickness. This is a result

7.1 Sequential Curing Process

163

of the applied parameterization scheme and the need of a minimal amount of


material in all regions in order to void singularities in the finite element run.
Laminates with constant thickness are easier to manufacture since the molds
do not have to follow the thickness distribution. For plane laminates, simple
plates can be used as forming components. Thus, the basic laminates are manufactured in a first curing cycle neglecting all the reinforcement layers. This
is schematically shown in Figure 7.2(a). The basic laminate is located between
two plates in order to have two identical surfaces. The curing process is carried out in a vacuum bag which is indicated with the green line. In the next
step, the reinforcement layers are applied to the basic laminate. Since it is already cured, it can be used as the forming component wherefore no additional
molds are needed. The reinforcement layers are attached symmetrically and
the compound is enclosed in a second vacuum bag which perfectly follows the
varying thickness distribution. While the reinforcement layers are cured in a
second temperature cycle, they simultaneously bond to the basic laminate. The
second curing step is illustrated in Figure 7.2(b). The critical point using the

(a) Curing cycle 1: basic laminate

(b) Curing cycle 2: reinforcement layers

Figure 7.2: Sequential curing process


sequential curing process may be the interlaminar shear strength between the
basic laminate and the reinforcement layers. Since a good bonding is essential
for a high quality composite, the shear strength has been investigated more
detailed. Song et al. [156] investigated the effect of different manufacturing
methods on the shear strength of single-lap bonded joints. They carried out
shear tests for joints which have been cured in two steps (what they are calling
co-bonding) and compared them to simultaneously cured joints and adhesive
bonded laminate joints. However, they always added an adhesive film to the
specimens which had been cured sequentially. From the authors perspective,
this is not necessary as long as the resin system of the cured and the uncured
parts are identical and the surface is prepared appropriately. To confirm that

164

Manufacturing-Related Aspects

assumption, a series of single-lap bonded joints were manufactured and tested


within the Semesters thesis of B. Schneuwly [148]. All specimens have the
dimension shown in Figure 7.3. Three types of specimen were tested:

Figure 7.3: Single-lap bonded joint specimen [148]

1. prepreg-prepreg (simultaneous curing)


2. prepreg-laminate (laminate surface made with peel ply)
3. prepreg-laminate (laminate surface polished)
The first type is used as benchmark while numbers two and three consider a configuration of sequential curing with different surface. The shear
strength s has been calculated by the fraction of failure force F and the bonding area A b which is indicated by the red rectangular. Five specimens of the
first type and six of each of type 2 and 3 have been tested. This is not too much,
but it gives an idea of the shear strength loss due to the sequential curing.
The results are shown in Table 7.1. The tests confirm that there is no signifSpecimen

1
2
3

s [MPa]

24.9
26.2
22.7

std( s ) [MPa]

3.1
2.7
3.5

Table 7.1: Single-lap bonded joint test results


icant strength loss. Specimen type 2 has a slightly increased shear strength
compared to the reference configuration. This may be due to the relatively low
number of specimens. However, the peel ply surface has been chosen due to its
rough surface which may result in a better bonding. If polishing the specimens,
the fibers are uncovered. This is not wanted since the bonding of reinforcements and basic laminate is done by the matrix material. Therefore, the shear
strength of the third specimen type is lower, but still in an acceptable range.
Even if this study is not extensive, it can be assumed that the shear strength

7.1 Sequential Curing Process

165

in the bonding layer is not critical.


All specimens for verification have been manufactured with a sequential curing process using the bonding method with peel ply surface (type 2). In order
to control the bonding quality visually, polished specimens have been prepared
within the Semesters thesis of B. Rentsch [140]. A photomicrograph of a typical specimen is shown in Figure 7.4. On the left side, the basic laminate with a

reinforcement
layers

basic laminate

45

-45

90

90

-45

45

90 -45

45

boundary layer
~0.25mm

Figure 7.4: Photomicrograph of a boundary layer [140]


layup of (0/45/-45/90)S with a single ply thickness of 0.25 mm can be recognized.
The overall laminate quality is good. There are no large voids or extremely
resin-rich regions recognizable and the fiber volume content is high. The reinforcement layers are attached on the right side. In between, the boundary
layer can be noticed. The gray area is nothing else than the spare resin from
the first curing cycle. The right side is waved which is caused by the rough
surface of the peel ply. This waviness can be reduced by polishing or sandblasting the surface before attaching the reinforcement layers. However, according
to the tests above, the interlaminar shear strength may be reduced. Finally,
it can be said that the sequential curing process ensures a sufficient bonding
between basic structure and reinforcements. For none of the tested specimens,
not even for the notched plates, delamination of the reinforcements has been
noticed.

166

Manufacturing-Related Aspects

7.2 Reinforcement Layer Manufacturing


Considering the design of the reinforcement layers in Section 6, the representation is still element-based which is of course not practicable for real structures. However, the contours are mainly continuous wherefore they can be
smoothened. Within this thesis, this has been done manually using a vector
graphic editor. Considering the first example of the eigenfrequency splitting in

(a) = 90

(b) = 45

(c) = 45

(d) = 0

Figure 7.5: Contour smoothening for cutting the reinforcement layers


Subsection 6.1.3, the contours may result as shown in Figure 7.5. The original
shape of the 0 -reinforcement layer shown in Figure 6.9(d) is rather scattered
wherefore it has been simplified according to the shape shown in Figure 7.5(d).
Slight simplifications are also made for the other layers. The contours shown
in Figure 7.5 can be directly interpreted by a textile cutting machine. It must
be mentioned here that the contour of the panel, namely the outer shape and

7.2 Reinforcement Layer Manufacturing

167

the cut-outs, are machined after the second curing cycle. Thus, all lines that
do not cover the finite element mesh can be chosen arbitrarily. Connecting the
regions that will build the reinforcements will later ease the application. Also
the outer contour of the reinforcement layers is chosen slightly bigger. It can
be taken as reference when applying the single reinforcements sequentially.
Additionally, the laminate boundaries are usually of bad quality. They are removed by cutting the panel to its final shape.
The layer application is carried out according to Figure 7.6. The single layers

45

-45

90

Figure 7.6: Positioning procedure for the reinforcement layers


can be aligned easily by the outer rectangular shape which is here indicated
by the transparent rectangular. Practically, the upper protection layer of the
PREPREG is kept until the layer has reached its final position. Of course, the
protection layer is cut as well by the cutting machine. However, it can be simply taped afterwards to guarantee global connectivity. The reinforcements are
attached on both sides in order to obtain a symmetric laminate. A front view

168

Manufacturing-Related Aspects

image of the stacked contours is given in Figure 7.7. It can be understood as


a thickness plot where dark areas indicate regions with many reinforcement
layers. The analogous plot of the finite-element-based contours is given in

Figure 7.7: Front view of the contours for cutting


Figure 7.8. Choosing the finite element mesh with an appropriate refinement,
the mapping of the contour can obviously be done rather accurately. After the
application of the reinforcement layer, the plate is enclosed in a vacuum bag.
As described above, the basic plate acts as the forming component. The vacuum
bag fits the locally varying thickness distribution. After curing, the outer shape
and the cut-outs of the panel are machined by water jet cutting. A front view of
the final panel is shown in Figure 7.9. A comparison of the physically realized
panel with the finite-element-based plot (Figure 7.8) emphasizes, that the obtained solutions can be produced quite precisely. More detailed pictures of the
panel are shown in Figure 7.10. The fiber orientations of the reinforcements
are clearly visible. Due to the separation foil which has been employed in the
second curing cycle, the reinforcements are covered with thin film of resin material which makes the surface shiny. Close to the reinforcement edges, sparse
resin material has diffused on the basic plate surface. This additional resin is
advantageous for a proper bonding. The matrix forms a epoxy bevel at the reinforcement edges which reduces the stress concentrations (see Section 4.5.3).
Even if the process has been defined to manufacture a few specimens for the

7.2 Reinforcement Layer Manufacturing

Figure 7.8: Front view of finite-element-based contours

Figure 7.9: Front view of the manufactured panel

169

170

Manufacturing-Related Aspects

Figure 7.10: Details of the manufactured panel


validation of the presented method, it might be interesting to employ it for a
higher quantity of parts. The precision of the contours can hardly be realized
using a LCM process.

7.3 Process Improvements

171

7.3 Process Improvements


The presented process is rather simple but the quality of the resulting structures is high. An improvement of the quality can basically only be reached if
the process is automated. Bringing the process to the next level, namely to an
automated production with the possibility of a mass production, there are some
drawbacks which will be discussed in the following. Primarily, the structure is
cured in two cycles, one for the basic structure and one for the reinforcements.
This requires the structure to be heated twice which is sub-optimal in terms of
energy consumption. Secondly, the entire process has been done manually so
far which is not practicable if the number of parts is increasing.
Within this thesis, basic considerations for further process development
have been made. However, the author wants to add that none of those have
been realized or tested. The Boeing Company started to manufacture large
scale structural parts of CFRP. They addressed the problem of how to compact
uncured plies on the surface of composites structures applying for a patent [17].
There, a flexible, portable and reusable compaction sheet is introduced which
is placed locally over the uncured plies. It is coupled with a vacuum source
which evacuates the vacuum chamber and draws the sheet against the plies.
The sheet allows temporarily and locally applying pressure to the plies instead
of load it externally or using a vacuum bag for the entire structure. Inspired
by that idea, some further considerations have been made in order to make
the process simpler and more economical. While the local reinforcements are
compacted to the basic structure, they can be cured simultaneously applying
heat. This could potentially be done taking advantage of simple heating blankets which are available on the market and often used in so-called repair tool
kits. However, heating and compaction can be done simultaneously using a
device as sketched in Figure 7.11 could be used. The device consists of a vac-

Figure 7.11: Compaction and curing device


uum chamber and a heating chamber which is separated by the green colored
membrane. The vacuum chamber is evacuated through the air tubes which are

172

Manufacturing-Related Aspects

indicated with the black arrows. Simultaneously, a heating medium circulates


trough the heating chamber and cures the local plies. If additional compaction
force is needed, the pressure inside the heating chamber can be increased. In
order to prevent the device from lifting off the structural part, the structural
part, a load is applied to the centered shaft. Since the heat is supplied by a
circular flow, the heat is applied immediately and no additional time to heat
the system is needed. The devices can be supplied with a central unit if many
parts are cured simultaneously. Such a second curing process would speed up
the process and save energy since the structure is only heated locally. Dependent on the temperature, todays matrix systems have curing times of about
5 to 10 minutes. However, the second curing cycle can be done in parallel for
many parts which reduces the cycle time and the process becomes economical.
The cutting of the reinforcements with cutting machines is state-of-the-art.
The step of bringing the patches to its final place is probably the most challenging part. However, an automated placement with robots is realizable. The
single reinforcement layers can be attached together in a first step and applied
to the basic structure in a second step.
The concept of having a basic structure and applying reinforcements in a second step might be interesting for mass production where each part may differ
slightly. Considering the automotive industry, each car is designed to the customer requirements. Different mass distribution or performance may result in
different loads. A future goal could be to not only provide specific equipment to
the customer but also design the structure to the resulting requirements. This
is an important step for pushing the light-weight design to its limit. The proposed sequential curing process could be an interesting approach to customize
the structural parts using local reinforcements. Whether such a system could
be profitable and basically, whether the application of the proposed locally reinforced solutions could be feasible cannot be answered.

Chapter 8

Validation
In this Chapter, a selection of application examples presented in Chapter 6 are
compared to experimental results in order to validate the simulations arising
from the GLM. In the first section, the experimental results of the panel, which
has been reported in Chapter 7, are compared to the simulation. The manufactured panel has been analyzed using a scanning vibrometer which is able
to measure the eigenfrequencies and visualize the corresponding eigenmodes.
The comparison shows a high degree of accordance.
In the second Section, the notched plate which has been reinforced for double
strength (see Section 6.4) is tested. A visual method (Digital Image Correlation) and an acoustic method (Acoustic Emission) have been used to detect
first matrix cracking. Even if first ply failure cannot be identified clearly, the
strength increase due to the local reinforcements can be confirmed.

8.1 Vibrating Panel


This section validates the method for eigenfrequency problems on the example of the vibrating panel introduced in Section 6.1.3. Goal of the optimization
was to split the second and the third eigenfrequency to generate a resonancefree band. The corresponding plate has been manufactured as described in
Section 7.1. The simulation results are validated by comparing the six first
eigenfrequencies and the corresponding eigenmodes with experimental measurements. The simulation of higher eigenfrequencies may be increasingly affected by the discretization of the model and thus differ from measurements.
The experimental measurements are made with a so-called scanning vibrometer. It is able to perform non-contact measurements and visualizations of

174

Validation

structural vibrations. The velocity of different pre-defined points is measured


by taking advantage of the Doppler effect [183]. With the velocity information and the coordinates of these points, the vibration modes can be visualized
equivalent to the modes of the finite element model. According to the simulation in Section 6.1, the panel is made of an unidirectional CFRP-PREPREG (see
Table A.1) with layup (0/45/-45/90)S . The panel has been placed horizontally
in a box attaching rubber cords on each corner as shown in Figure 8.1. This
comes close to a free supported configuration. In order to make the model as

rubber cords

copper coil

Figure 8.1: Test set-up for vibration measurements [140]


realistic as possible, the rubber cords have been modeled with spring elements
as mentioned already in Section 6.1. The tensile moduli of the rubber cord has
been determined with a tensile test and linearized in the range of the occurring
loads. As excitation source, a copper coil is placed at the center point below the
panel. As counterpart, a small piece of steel is attached to the panel which is
considered in the simulation with the 10 g centered mass. The panel surface
is covered with silver color in order to improve the reflection of the laser beam
which is essential for the quality of the measurements.

8.1 Vibrating Panel

8.1.1

175

Panel with Basic Laminate

Amplitude [m/s]

In a first series of measurements, the six lowest eigenmodes of a panel without


reinforcements are considered. These results are used as reference to demonstrate the frequency shifting caused by the reinforcements. In case that the
frequencies and the modes of the simulation match the experimental data, it
can be assumed that the basic finite element model, e.g. geometric representation, boundary conditions and material data, are adequate and the model can
be used for the optimization. In case of a strong discrepancy between model
and experiments, the model needs to be tuned. The six lowest eigenfrequencies of the basic panel are expected to be in a range between 40 Hz and 200 Hz.
The panel is thus excited with a sweep in this range. Figure 8.2 shows the veMeasurement
Simulation

0
10

1
10

2
10

40

60

80

100

120
Frequency [Hz]

140

160

180

200

Figure 8.2: Frequency band of the unreinforced plate


locity amplitude in the frequency domain between 40 Hz and 200 Hz in a logarithmic scale. The eigenfrequencies can be identified by the amplitude peaks.
The maximal amplitude value is emphasized with the red lines. For these six
frequencies the amplitudes are combined corresponding to the measurement
points mesh in order to create a visualization of the eigenmodes. The predicted
eigenfrequencies of the simulation are additionally indicated with green lines.
According to Figure 8.2, measurement and simulation results match each other
well. A numerical comparison between simulative and experimental results is

simulation
experiment
f /f sim

[Hz]
[Hz]
[%]

f1

f2

f3

f4

f5

f6

47.45
47.50
+0.10

76.63
76.25
-0.49

82.00
82.81
+0.99

119.61
118.4
-0.98

137.03
135.00
-1.48

188.14
185.00
-1.67

Table 8.1: Basic panel: result summary

176

Validation

given in Table 8.1. It confirms that the results are of high accordance. The
first four modes deviate by less than 1% and the others by less than 2%. This
is much more accurate than expected since the finite element model contains
simplifications. The eigenmodes for both, experimental and simulative results,
are contrasted in Figures D.1 to D.6 in Appendix D.1. Again, the agreement between simulation and experiment is high. With the exception of mode 4, which
is the one with lowest amplitude, the modes are almost identical. Based on
these results, it can be assumed that the finite element model represents the
real panel very realistically. Therefore, the model has not been changed anymore for the optimization. The author wants to emphasize that all material
parameters used within the model arise from tensile tests and no parameter
tuning has been made.

8.1.2 Panel with Local Reinforcements

Amplitude [m/s]

The optimization and the manufacturing of the panel are discussed in Sections 6.1.3 and 7.1. The same measurements have consequently to be taken
with the locally reinforced panel. The test set-up is equivalent to the one
for the experiments of the basic panel. According to the prediction (see Table 6.3), the reinforcements should generate a resonance-free band between
71.44 Hz and 267.52 Hz. Thus, the sweep for excitation of the panel has been
extended to 300 Hz. The corresponding plot of the velocity amplitudes is shown
in Figure 8.3. As before, the amplitude peaks of measurement and simulation
Measurement
Simulation

0
10

1
10

2
10

50

100

150
Frequency [Hz]

200

250

300

Figure 8.3: Frequency band of the reinforced plate


are emphasized with red and green lines. The sixth eigenmode which should be
at 267.52 Hz could not have be detected in the experiment. It is assumed that
this mode is very weak. Also, a measurement where the copper coil has been
moved in order to eliminate the possibility of an excitation at a node line did

8.1 Vibrating Panel

177

not help to locate the mode. The numerical results are summarized in Table
8.2. Excepting the third mode, the relative deviation is below 2%. The modes

simulation
experiment
f /f sim

[Hz]
[Hz]
[%]

f1

f2

f3

f4

f5

f6

71.40
70.00
-1.96

79.06
78.13
-1.17

208.68
200.00
-4.16

211.26
207.50
-1.78

225.65
223.13
-1.12

267.52
-

Table 8.2: Reinforced panel: result summary


are contrasted in Figures D.7 to D.11 in Appendix D.2.
Also here, the experimental and the simulative modes are almost identical. Only mode 4 shows a slightly different shape. Generally, the deviation
between experiments and simulation is bigger for the reinforced panel. This
can be simply explained by the fact that the finite element model discretizes
the contours of the reinforcement layers. Moreover, there are some discrepancies arising from the contour smoothening and the manufacturing process.
The cutting and the placement of the reinforcements always imply some imprecisions. However, recalling the goal of the optimization, which is to split
the second and the third eigenfrequency in order to clear a frequency band,
the agreement between the results is still remarkably high. The experiments
confirm the existence of a frequency band between 80 Hz and 200 Hz where no
eigenfrequencies occur. Thus, the experiments demonstrate that the presented
GLM is able to specifically move the eigenfrequencies of a structure.

178

Validation

8.2 Notched Plate


The second example for validating the presented GLM is the notched plate introduced in Section 6.4. Goal of the optimization is to double the first ply failure (FPF) load according to the Tsai-Hill criterion. The resulting reinforcement
layers are shown in Figure 6.22. Specimens have been manufactured using the
sequential curing process presented in Section 7. They are made of an unidirectional CFRP-laminate (see Table A.2) with a basic layup of (0/+45/-45)S . For
this configuration, only in-plane loads and no bending occur. The classical laminate theory homogenizes the material properties through the thickness and the
membrane stiffness becomes invariant in terms of stacking position. Therefore,
the stacking of the reinforcement layers can be chosen arbitrarily without disregarding the simulation. Based on that, four different types of specimens have
been manufactured. A schematic overview is given in Figure 8.4. Type A is the

(a) type A

(b) type B

(c) type C

(d) type D

Figure 8.4: Specimen types of the notched plate


reference configuration made of the basic laminate (0/+45/-45)S with no reinforcements. It is used to quantify the strength increase due to the reinforcements. Also type B, which consist of the double basic laminate (0/+45/-45)2S ,
is used as a reference. It is expected to have twice the structural strength of
type A. However, not only strength but also its mass is doubled. Type B will
help to identify strength reduction due to local stress concentrations caused
by the reinforcements. Types C and D are specimens having local reinforcements. Figure 8.4 schematically shows the different stacking where the dark
layers represent the reinforcements. For both, the reinforcement layer shapes
are identical corresponding to the simulation results obtained in 6.4. However,
for type C the reinforcements are attached externally using the sequential curing process. For type D, the reinforcements have been placed between the 45 and the 0 -layer. Analogously to the reinforcements, the 0 -layer has been attached after the first curing cycle. Initially, the reinforcements for type C were
expected to peel away due to interlaminar stresses. The risk of interlaminar
failure could be eased using the specimens of type D. However, for none of the
C-specimens, failure at the interfaces between basic and reinforcement laminate could be noticed. The number of manufactured specimens is 5 for type

8.2 Notched Plate

179

A and B and 10 for type C and D. According to the simulation, the reinforcements have been generated to double the FPF-load. Excepting unidirectional
laminates, FPF is usually caused by matrix cracking. For the validation here,
a method for identification of matrix cracking is needed. However, there exists
no standardized method for it. Here, two complementary measurement techniques have been applied. Digital Image Correlation (DIC) has been used as
visual method and Acoustic Emission (AE) measurements as acoustic method.
The application of two techniques based on different physical principles is expected to give a more complete picture of the evolving damage. Both methods
and their results are presented in the next Subsections. The specimens have
been loaded at constant cross-head speed in a uniaxial tensile machine until
ultimate failure. The test set-up with the camera for DIC and the sensors for
the AE-measurements is illustrated in Figure 8.5.

clamps

Camera for
DIC

AE-sensors

specimen

Figure 8.5: Tensile test set-up with camera for the DIC and acoustic sensors
for the AE-measurements

180

Validation

8.2.1 First Ply Failure


Digital Image Correlation
Digital Image Correlation (DIC) is an optical method to measure deformation
or strain on the surface of an object. It is widely used in science and engineering due to its simplicity in application. A detailed explanation of the method
is given in [27, 18]. A review of research work using DIC for displacement
and strain measurements is given by Pan et al. [123]. The DIC is used to
compare digital images taken at different times and loads, respectively, and
to determine the changes of deformation between them. The out-of-plane deformation of the considered specimens is insignificant, wherefore a 2D-DIC,
which only requires a single camera, is feasible. However, there exist also 3DDIC methods for the measurement of more complex deformation fields which
need multiple cameras. Each pixel is typically converted into an eight-bit number where 0 represents white and 255 black. The correlation function S (after Bruck et al. [18]), which measures how well a subset of correlation points
match, has the form
*


u u v v
(F(x, y) G(x, y))
(8.1)
= 1  
,
, ,
S x, y, u, v,
 *

*
x y x y
F(x, y)2
G(x, y)2
Thereby, F(x, y) are the gray level values of the reference image at point x, y
and G(x, y) for the second image at point x*,y*. The coordinates x, y and
x, y are connected by the deformation.
x = x + u +

y = y + v +

u
x
v
x

x +

x +

(8.2)

(8.3)

y
v
y

The variables u and v are the displacements for the coordinates x and y and
x and y are the distances from the subset center to point x, y. A subset
of correlation points is manually defined by the user. Minimizing the correlation function S provides the new coordinates x, y or the displacements u, v,
respectively. Deformation or the strains can be calculated with these results.
Lomov et al. [96, 68] take advantage of DIC to identify damage initiation in
textile composites. They consider the standard deviation of the strain or of its
relative error, respectively. Damage is detected if some of the values come out
of the probable value range. The application of this method may become problematic if the considered strain-field is not homogeneous as it is the case at the
hole of the notched plate. Here, the DIC is applied in a similar way. It is basically used to find non-linearities in the displacement path of the correlation

8.2 Notched Plate

181

points. As long as the specimen has no cracks, the load-displacement relation is


assumed to be linear. The displacement of the correlation points increases thus
proportional with the increasing force. If first cracks occur, the stiffness of the
specimen is changed. Consequently, the linear load-displacement relation is
disturbed. The detection of the stiffness change can be used to identify damage
events. However, the load for which a non-linearity in the load-displacement
relation is detected cannot directly be interpreted as FPF-load. Optically detectable damage is the consequence of a number of sub-critical damages which
occur already at lower load levels. Moreover, only a small section of the specimen is monitored and only mechanisms changing the surface displacement
field are detected. Nevertheless, a correlation between FPF-load and the load
where a first non-linearity in the load-displacement relation is detected is expected. Thus, for specimen types B, C and D, drifting of the correlation points
should be noticed for twice the load as for specimen type A.
The simulation identifies the maximal material load at the hole edge of
the 45 -layer. Figure 8.6(b) shows that they are slightly shifted to the right
side (and left for the symmetric side) of the vertical symmetry line. The first

FImax

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

(a) = 45

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0.05

(b) = 45

0.15

0.25

0.35

0.45

0.55

(c) = 0

Figure 8.6: Failure indices (Tsai-Hill) for the unreinforced notched plate
optically detectable damage events will occur in the high loaded areas according to Figure 8.6. Even if the first cracks are in the 45 -layer, which is not
directly observable, they will have a detectable impact on the stiffness. A digital camera with a resolution of 1624 x 1234 pixels has been placed to record
the strains in these areas. The recorded section has a dimension of roughly
14.2 mm x 10.8 mm. An impression of the dimensions is given in Figure 8.7.
The exact dimension is not important since the absolute values of the displacement are not of interest. The smaller the section considered, the higher the

182

Validation

Figure 8.7: Recorded section for the DIC


displacement resolution. However, due to the slipping of the specimen, the
considered section should be chosen big enough to ensure that the tracked
points do not leave the image during the test. Unfortunately, the section of
interest left the observed area for one specimen each of type A and C, and two
of type D which, therefore, have been eliminated for the evaluation. Images
are taken with a sampling rate of 2 Hz. Additionally, the load signal of the
tensile test machine is recorded and synchronized to link the load with the corresponding strain field. The experiments have been made taking advantage
of a self-implemented LabVIEW 1 environment. A grid of 36 correlation points
has been defined in a square formation on a tangential line in load direction
as shown in Figure 8.8(a). Each point is detected 4 times by moving the initial
coordinate slightly. This enables a filtering of points that step out of line due to
problems in the correlation process which typically happens if the contrast of
the image is insufficient. In order to ensure that the correlation point cluster is
positioned equally for all specimens, the hole has been detected automatically
with image recognition and coordinates have been set relatively to it.
Figure 8.8(b) shows a configuration of correlation points for a loaded speci1 LabVIEW is a system design software from National Instruments to create and deploy measurement and control systems

8.2 Notched Plate

183

(a) initial configuration

{
{

1
2 {
3
(b) loaded configuration

Figure 8.8: Reference (black) and displaced (white) correlation points for the
initial and a loaded configuration

men. The cracks initiated by the hole are clearly visible. The cluster has been
split into three sub-clusters indicated with 1 , 2 and 3 . The 2 points are
distributed almost randomly. Since they are directly located on the crack, the
image has changed so significantly that a tracking of these points is not possible anymore. The points of 1 and 3 are still arranged in a block. Their relative displacement in load direction is significantly different. The corresponding
load-displacement relation is shown in Figure 8.9(a). After a fast increase of
the displacements which is caused by a slight slipping of the specimen in the
clamping jaws of the tensile test machine, the displacements increase almost
linearly. The slopes of the load-displacement curves are different for each correlation point since the relative displacements between the points are non-zero.
For higher loads, the displacement values of some correlation points become
non-linear before the specimen finally fails. For these load ranges, huge cracks
are already visible wherefore the DIC is not able to track the points anymore
which leads to a kind of random displacement value. However, in the range
where FPF is expected, it seems that the load-displacement relation is approximately linear and no extensive non-linearities can be noticed. Taking the first
derivative of the displacement with respect to the force provides a measure
of stiffness, or compliance respectively, which is shown in Figure 8.9(b). The
compliance is not constant as it is expected from a linear behavior. The reason
for that might still arise from the clamping or from a stretching of the fibers
which are not ideally straight in the initial configuration. However, a bifurca-

Validation

45

45

40

40

35

35

30

30

Load [kN]

Load [kN]

184

25
20

25

15

15

10

10

100

200
300
Displacement [pix]

400

(a) Load-displacement plot

500

20

7.5

8
8.5
9
Compliance [pix/kN]

9.5

10

(b) Load-compliance plot

Figure 8.9: Load-displacement and load-compliance plot for all correlation


points of one particular specimen
tion of the stiffness is clearly visible. Basically, the curves are split into three
groups which are of course connected to the three groups in Figure 8.8(b). This
bifurcation is already detectable in the plots before it can be noticed in the correlation picture. The compliance values which are assumed to be constant are
different for the correlation points. Different values can be mitigated taking
the second derivative of the displacements with respect to the load. Assuming that the load-displacement relation is linear, the second derivative must be
equal to zero for all correlation points. It can be understood as the acceleration
of the correlation points dependent on the applied load. As long as the stiffness
is constant, the points move with constant velocity and consequently the acceleration is zero. However, a stiffness change provokes a peak which can be used
to identify damage events. Since results arising from DIC are rather noisy, the
displacements are filtered with a Butterworth-lowpass filter before taking the
derivatives. Figure 8.10(a) shows the deviation of the second derivative of the
displacements corresponding to equation (8.4).
di =

2 u i
f 2

n 2 u
1 
j
n j =1 f 2

(8.4)

Since its mean value is zero, it expresses the disturbance of the linear behavior. Within the low load range, the values are all close to zero. With increasing
loads, the values become non-zero which indicates a material non-linearity or
first cracking, respectively. A load value which is taken as FPF-load is determined using the standard deviation of the correlation point accelerations
shown in Figure 8.10(b). FPF is assumed if the function value exceeds a pre-

185

45

45

40

40

35

35

30

30

Load [kN]

Load [kN]

8.2 Notched Plate

25
20

25
20

15

15

10

10

0
100

50

0
50
Acceleration [pix/kN2]

(a)
Correlation
deviations d i

point

100

10

20
30
40
std(Accelertation) [pix/kN2]

50

acceleration (b) Standard deviation of the correlation


point accelerations

Figure 8.10: Plots of the correlation point accelerations


defined threshold. This threshold is set manually and there is no physical
meaning. Since the goal here is to verify the load increase for the different
specimen types, it is important to take the same value for all specimens. For
the evaluation here, the value has been set to 1. Taking a threshold of 2, the
averaged loads are increased by a percentile range of 10% to 30% dependent
on the speciment type. This emphasizes that the loads are only feasible for
relative comparisons between the specimen types.
A summary of the loads at which cracks have been detected is given in Table 8.3. The first and the second row give the average failure load and the
Specimen type
F D IC [kN]
std(FD IC ) [kN]
F D IC [%A]
F D IC [%FFPF ]

A
10.01
1.23
-10

B
23.70
2.70
+137
-6

C
19.63
2.37
+96
-12

D
15.96
1.44
+59
-28

Table 8.3: Load at occurrence of first visual crack


corresponding standard deviation. The third row provides the load increase
with respect to the failure load of specimen type A. Considering the simulation, the load increase should be +100% for specimen types B,C and D. For
specimen type B, which has a double layup, the load increase is even higher
than expected (+137%). The load increase for the locally reinforced specimens
is less than expected. The fourth row gives the deviation from the predicted
value of the simulation. One specimen each of type A and C, and two of type D

186

Validation

are eliminated since the area in which first cracks occur has left the considered
section. Additionally, a third specimen of type D is eliminated due to slipping
in the clamps before breaking.

Acoustic Emission Measurement


The presented method for the early optical detection damages, which are the
result of sub-critical damage accumulation, with Digital Image Correlation is
novel and thus not generally accepted. Even if the method does not directly
identify FPF, the resulting loads are expected to correlate with the actual
FPF-loads. Visual damage is generally detected at higher load levels. Supplementary monitoring with Acoustic Emission (AE), which is a method that is
sensitive to microscopic damage mechanisms and events, was performed for all
tensile tests with the aim of providing an independent measurement of damage
and specifically of FPF for a comparison with the values obtained from 2D-DIC.
Both measurements have been made on the same specimens simultaneously
which allows a direct comparison of the respective results. A few publications
report the application of AE-measurements for the detection of sub-critical
damage. Experimental predictions of FPF-strength of laminated composite
plates are made by Kam and Lai [74]. Pahr and Rammerstorfer [122] use AE to
obtain FPF-strength values of woven fabric laminates. Sihn et al. [154] investigate the strength of thin-ply laminates by measuring AE on notched plates.
The AE-technique is used by Cesari et al. [23] to detect damage progression
in circular composite plates. An experimental analysis of weave composites
using AE is presented by Daggumati et al. [31]. Chang and Chiang [24] used
AE signal energy rate for the determination of FPF in GFRP plates. Basically,
AE-measurements are complex and require a fundamental understanding of
the measurement equipment. Thus, they have been carried out and evaluated
by Dr. A. Brunner from the Swiss Federal Laboratories for Materials Science
and Technology. Four AE-sensors, 2 of type SE-150M and 2 of type SE-45H
from Dunegan Engineering Inc., have been clamped on the specimens according to Figure 8.5. Additionally, two two SE-150M sensors have been used as
guard sensors on the clamps for eliminating AE signals originating from the
load introduction. The data have been recorded with the commercial equipment AMSY-5 from Vallen Systeme GmbH. The sampling rate has been set to
10 MHz, the threshold to 40 dBAE , the rearm time to 3.2 ms and the signals
were frequency filtered with a band-pass between 25 and 850 kHz and amplified by 34 dB. The displacement and load data from the tensile machine were
synchronously recorded with a rate of 4 Hz. The data analysis was performed
with the equipment specific software AEVisual from Vallen Systeme GmbH.

8.2 Notched Plate

187

It is difficult to assign specific AE-signals to damage events. Considering


the literature cited above, damage is empirically assessed based on AE-activity
and AE-intensity as a function of load. The results presented below are from
an analysis of AE intensity, e.g. AE signal energy rate versus time or load,
analogous to that by Chang and Chiang [24]. The signal energy is measured
in energy-units [e.u.] which are equivalent to 1014 V2 s. Figure 8.11 shows an
example of specimen type D (AE sensors SE-45H) where the AE energy rate
(AE signal energy per 0.1 second) and the load are shown versus time both, in
a linear (Figure 8.11(a)) and a logarithmic scale (Figure 8.11(b)). As criterion
x 10

45
40

Energy [eu]

12.5

35

10

30
25

7.5

20

15
10

2.5
0
0

Load [kN]

15

5
50

100
Time [s]

150

0
200

10

45

10

40

10

35

10

30

10

10

10

10

10

25
20

Load [kN]

Energy [eu]

(a) Full scale plot yielding three major energy bursts at


about 110, 130 and 170 seconds

15
10
5
0

50

100
Time [s]

150

0
200

(b) Plot in logarithmic scale yielding major AE energy bursts


already between 100 and 110 seconds with threshold (dashed
line) of 5x105 e.u.

Figure 8.11: AE energy rate (per 0.1 second) for one type D specimen from the
data recorded with the SE-45H type sensors
for the first major energy rise as identification for FPF[24], an AE energy rate

188

Validation

threshold of 5x105 e.u. per 0.1 seconds per sensor type has been used. Some
increased AE activity has been detected at the initial loading which is assumed
to be related to the slipping of the specimens though the clamps wherefore
the FPF loads are further filtered. The failure strain of unidirectional CFRP
materials is roughly 1.8% in fiber direction and 0.5% perpendicular to it. With
an additional margin, it can be assumed that no FPF-events may occur below
a strain of 0.3%. Taking advantage of the FEM model, corresponding loads
of 3.52 kN for specimen type A and 7.04 kN for specimens B, C and D were
determined and FPF loads based on the AE measurements below those limits
were replaced by the corresponding next higher AE signal energy rise.
Table 8.4 summarizes the results of the experimentally determined FPFload for the four specimen types. On average, a significant load increase for
Specimen type
F AE [kN]
std(FD IC ) [kN]
F AE [%A]
F AE [%FFPF ]

A
8.15
2.35
-27

B
16.35
4.14
+101
-27

C
13.98
3.54
+72
-37

D
14.81
3.96
+82
-36

Table 8.4: Tensile load at first occurrence of 5x105 e.u./0.1s


specimens B, C and D can be noticed. In contrast to the DIC-results, the average load of type B is very close to the expected 100%. The measured loads
of the locally reinforced specimens C and D are again below the values predicted by the simulation. In general, the standard deviation is higher than
for the DIC-measurements. One specimen of type A has been eliminated due
to rather low AE intensity and one specimen each of type B and C have been
eliminated due to detection of increased AE signal energy at unrealistic low
loads as described above. Also here, the specimen of type D, which slipped in
the clamps is eliminated.
Correlation and Interpretation of DIC and AE results
In order to see whether the experimentally determined loads for FPF via the
two different methods agree, the data are correlated in Figure 8.12. The loads
at which DIC detects cracks are plotted versus the loads at which the AE signal
energy rate exceeds 5x105 e.u. per 0.1 seconds. Only specimens are plotted for
which both measurement methods provided a valid load value. Consequently,
the mean value of the points in Figure 8.12 do not directly correspond to the
one given in Tables 8.3 and 8.4. The dashed diagonal indicates where the loads
for FPF from the two methods would be equal. Figure 8.12, however shows that

8.2 Notched Plate

189

A
B
C
D

Load at First Optically


Detectable Damage [kN]

25

20

15
10

5
10
15
20
5
Load at First Energy 5x10 e.u./0.1s [kN]

25

Figure 8.12: Comparison for DIC- and AE-results for single specimens

most points lie above this diagonal. This is consistent with the fact that volumetric AE measurements are more sensitive to damage inside the specimens
than surface DIC. Visually detected cracks hence cannot be directly correlated
with AE events. The scatter of the individual data points for the different specimen types as well as for the two methods is large. The correlation coefficient
for all specimens is 0.44 which implies a very weak correlation. Material imperfections and the non-trivial manufacturing process quite likely are the main
reasons for this. Moreover, the optical DIC method considers only a limited section on the surface of the specimen while the AE measurements are sensitive
to events in the entire specimen.
The spreading of the points of the particular groups for both measurement
methods is high. Material imperfections and the non-trivial manufacturing
process are the main reasons for that. On a statistical basis, the analysis looks
much better. Figure 8.13 shows the same plot but this time with the average
loads per specimen type and test method. Consider that the averaged loads
contain all valid load values from each measurement method individually corresponding to Tables 8.3 and 8.4. Again, the data from AE show a more sensitive behavior and lower FPF loads. However, the correlation coefficient for the
average data is now 0.91 which implies a strong correlation between the two
methods. The solid black line in Figure 8.13 represents the linear regression
of all four points. The results of the locally reinforced specimens C and D behave contradictory. While the DIC-results provide a higher FPF-assumption
for specimens C, the AE-measurements predict higher loads for specimens D.
A possible explanation of this contradiction could be the different surface qual-

190

Validation

A
B
C
D

Averaage Load at First Optically


Detectable Damage [kN]

25

20

15

10

5
10
15
20
5
Average Load at First Energy 5x10 e.u./0.1s [kN]

25

Figure 8.13: Comparison for DIC- and AE-results for averaged specimens

ities of specimens C and D which influences both of the applied measurement


methods. While the surface of type C is smooth and matte due to the peel ply,
the surface of the D specimens is relatively rough and shiny which is caused
by the roving bundles of the outer 0 -layer. The differences in the visual properties may effect the detection of failure events using DIC. On the other hand,
the surface roughness has an impact on the AE-measurements since the coupling of the acoustic sensors is different. The incoming energy for an equivalent
event may be different which influences the determined load.
Especially for the DIC measurements, the strength increase of specimens
B compared to specimens A is more than the expected 100%. A reason for
that may be the distribution of the 0 -layers through the thickness. Related
investigations on the notched strength of composite have been made by several
authors [55, 48, 184]. Primarily, they consider ultimate strength and not FPF.
However, Green [48] presumes that 0 -layers impede the propagation of subcritical damages. While specimens A have only top 0 -layers, specimens B have
two more of them inside the laminate. Consequently, the propagation of these
sub-critical damage is more suppressed for specimens of type B which may lead
to a detection of damage at higher loads, especially when using a visual surface
method such as the DIC method.
Considering specimens C and D, the FPF-load has been increased significantly independent of the measurement method. However, the predicted increase of 100% is not reached (see Tables 8.3 and 8.4). This can be explained
with the stress concentrations caused by the stiffness inhomogeneity due to
the local reinforcements which are not covered by the applied shell elements

8.2 Notched Plate

191

(see Section 4.5.3). The 45 -layers are relatively small and located around
the critical point. The stiffness inhomogeneities may also influence the critical
stresses which may cause first matrix cracking at lower load levels. However,
the strength increase of the locally reinforced specimens is in a range of 59%
to 82% and thus still high compared to the additional mass of 6%.
Both measurement methods detect damage events at lower loads than predicted by the FEM-simulation. Considering Tables 8.3 and 8.4, the deviation
for the unreinforced specimens A and B is up to 30%. On one hand, this can
be attributed to the free edge effects at the hole edge which may add a significant contribution to the stresses. A detailed overview of investigations on
free-edge effects is provided by Mittelstedt and Becker [114]. The applied shell
finite elements only consider in-plane stresses wherefore these effects cannot
be covered. Regarding to the simulation, first damage is matrix cracking at the
hole edge due to a combination of shear loads and normal loads perpendicular
to the fiber direction. A slight overestimation of the shear strength S and the
tensile strength perpendicular to the fiber Yt may contribute additionally to
the error. A correction is not done here since the design of the reinforcements
is dependent on the strength values through the Pseudo Strength Function. An
additional error source may be the chosen failure criterion itself whose physical
basis is rather weak. However, the goal of the experiments is to demonstrate
the strength differences between the specimen types and not to tune the simulation with experimental data.

8.2.2

Ultimate Failure Load

The ultimate failure load is the load at which the specimen completely breaks
and no force can be carried anymore. Even if it was not directly the objective
to increase this load, it is increased as a consequence of the optimized first ply
failure load. The ultimate failure loads are discussed with the values in Table 8.5. The first line summarizes the average failure loads for the different
Specimen type

F ul t [kN]
std(F ul t ) [kN]
F ul t [%A]
F ul t [%B]
F ul t [%FFPF ]

22.79
2.67
-46.73
+104.64

42.78
2.59
+87.73
+92.09

37.21
2.03
+63.27
-13.02
+67.06

38.96
4.72
+70.98
-8.93
+74.94

Table 8.5: Ultimate failure load result table


specimen types. The sample size of types A and B are 5, 10 of type C and 8

192

Validation

of type D. Two specimens of D have to be eliminated since they failed at the


support plates or slipped through before breaking. The second row gives the
standard deviation of the sample. Obviously, the spreading of type D is significantly higher than that of the other types. The manufacturing process of
these specimens was more complicated which may have caused a bigger variance in the specimen quality. In the third row, the strength increase of the
specimens B, C and D with respect to the A specimens is given. The strength
increase considering the finite element model should theoretically be 100 %.
However, the B specimen has only an increase of 87.73%. Recalling again the
investigations [55, 48, 184] on the notched strength of quasi-isotropic materials, they all arrive at the conclusion that the notched strength decreases with
increasing laminate thickness. For quasi-isotropic laminates, they all arrive
at the conclusion that the notched strength decreases with increasing laminate thickness. Green et al. [48] attribute the failure stress reduction to
the distribution of the 0 -layers through the thickness. These plies stop the
propagation of sub-critical damages, which cause a local stress relief, confining them to the outer sub-laminates. The missing stress relief results in a
reduction of the failure stress. In contrast, Harris and Morris [55] investigated the notched strength for (0/45)S laminates as utilized here and came
to the conclusion that it increases with increasing thickness. The strength
increase is based on a change in the interaction of different fracture modes.
While the thin laminate fails as a result of an uncoupling mechanism caused
by delaminations at the 45 -interface and fiber breaking of the 0 -layer only,
the thicker laminate showed fiber breaking in all layers. However, the failure
modes of all tested specimens within this thesis were identical independent of
the laminate thickness. A typical fracture mode is shown in Figure 8.14. Corresponding to the failure mode of the thin specimens above, only the fibers of the
0 -layers failed as a consequence of the delamination in the 45 -interfaces.
Consequently, an increase of the strength due to a change in the interaction
of different failure modes cannot be expected. On the contrary, the effect observed for quasi-isotropic laminates, namely that thicker specimens have less
sub-critical damages which reduce the stress concentrations, could be responsible for the strength reduction of type B. However, a concluding statement of
that phenomenon cannot be given here. The strength improvement of the locally reinforced specimens (type C and D) is 63.27% and 70.98%, respectively.
Here, the stress concentrations due to the local reinforcements are playing a
role again. Independent of the discussion above, the strength reduction due to
the thicker laminate has to be taken into account as well since all specimens
fail at regions with double laminate thickness. From that point of view, the
strength of specimens C and D must basically be compared to the specimens of

8.3 Discussion

193

0-layer fiber cracking

45-layer delamination

Figure 8.14: Typical failure mode of the notched plate

type B to mitigate that effect. Considering row 4, the locally reinforced specimens come close to the strength of specimens B. This is remarkable recalling
that they have only 53% of the mass of B. The last row compares the ultimate
failure load to the corresponding predicted FPF-load. The failure loads are all
between 75% and 105% higher. This is reasonable since the FPF criterion is, at
least in this case, a rough estimation for matrix cracking. When dimensioning
a structure for FPF, there is still a huge margin until it fails completely.

8.3 Discussion
Two examples for the validation of the GLM have been presented within this
chapter. The agreement between simulation and experiments of the vibrating plate is remarkable. The GLM locally reinforces the quasi-isotropic panel
which leads to a local anisotropy giving it the outstanding structural properties. It is clearly demonstrated that the method is able to specifically reinforce
structures to a desired objective. The fact that the frequencies are not only
maximized, which is always achieved to some degree when adding material
to the structure, but that some are maximized while others are minimized
proves that the resulting design is not a random outcome. The dynamic behavior of structures is governed by the stiffness and the mass distribution.
Consequently, it can be assumed that the method obtains likewise good results
considering stiffness, or compliance problems respectively, as well as buckling
problems since they are only dependent on the stiffness properties of the structure.

194

Validation

Performing an optimization with respect to strength as for example the


notched plate additional material parameters, namely the material strengths,
have to be taken into account. In general, the result spreading of strength
tensile strength tests is much higher than for stiffness measurements. The
strength properties of laminated composites are dependent on the semifinished material quality, the curing and the manufacturing process. In contrast to the stiffness properties, strength properties can be strongly influenced
by local weakening. The result spreading, which can also be seen in the tensile
tests of the notched plate, is one reason why the accordance between simulation
and experiments is less than for the vibrating panel. However, more reasons
are responsible for the discrepancy. The chosen failure criterion of Tsai-Hill is
a rough estimate for FPF which is in this case matrix cracking. Its application
is problematic firstly since matrix cracking is not easy to identify and secondly
since the experimental determination of the required strength parameters is
not straight forward and there is a large spreading as well. Additionally, the
effects of the stress concentrations caused by the local reinforcements (see Section 4.5.3) are playing a decisive role. Dependent on the manufacturing process, the stress concentrations may vary significantly between the specimens.
Despite all the mentioned experimental uncertainties, the experimental results confirm a significant increase of the strength due to local reinforcements.
Considering all the error sources, the agreement between predictions and measurements is still in a reasonable range. Table 8.6 summarizes the strengthto-mass ratios calculated with the FPF load arising from DIC (D IC ), AE ( AE )
and with the ultimate load ( ul t ). Obviously, the ratios for both loads, FPF
Specimen type
m [g]

A
22.61

B
45.21

C
23.96

D
23.96

FD IC [kN]
D IC [kN/g]

10.01
0.44

23.70
0.52

19.63
0.82

15.96
0.67

F AE [kN]
 AE [kN/g]

8.15
0.36

16.35
0.36

13.98
0.58

14.81
0.62

Ful t [kN]
ul t [kN/g]

22.79
1.01

42.78
0.95

37.21
1.55

38.96
1.63

Table 8.6: Strength-to-mass ratios


and ultimate, have highest values for the locally reinforced specimens C and
D. This emphasizes that the potential of composite materials is only exploited
if the anisotropic material properties are utilized, e.g. by the application of local reinforcements. Finally it can be said that the two examples validated the
experimental results and the power of the underlying method.

Chapter 9

Discussion and Conclusion


This chapter contains a discussion on the presented work. Primarily, the presented method for the generation of local reinforcement is discussed. Considerations on the local anisotropy of the solutions caused by the reinforcement
layers are given. The thesis is completed with an outlook of potential further
research.

9.1 Discussion
The presented parameterization scheme including virtual ghost layers, together with the highly accurate semi-analytical calculation of sensitivities of
chosen objectives with respect to changes of layer thickness values, is the essential part of a strategy for adding local reinforcements to a given composite
structure in order to obtain high performance at very low weight. Alternatively, the method can also remove material where its effect on the efficiency of
the structural performance is negative. The power and the general applicability of the GLM are demonstrated by enhancing the method for fundamentally
different objectives. The method has then been applied to respective concrete
design problems, such as:
Minimal Compliance
Harmonic vibrations
Nodal displacement
Stability
Strength

196

Discussion and Conclusion

The method provides great freedom in defining and realizing objectives. It is


shown that it cannot only maximize the natural frequencies, but also shift the
eigenfrequencies of a set of chosen eigenmodes in order to change the distance
between them. Obviously, the resulting structures corresponding to these objectives are difficult to find by intuition which emphasizes the value of the
method for practical design. More design solutions with unexpected reinforcement patterns but outstanding structural performances are presented for problems of stability and nodal displacements. Structural strength is limited by a
history of damage events which is extremely difficult to model and implies a
huge numerical effort. Therefore the here considered improvement of strength
addresses only the increase of FPF load. The quantitative verification by comparison with physical measurements is very difficult. However, it is shown
that the objective of achieving higher FPF loads can lead to design solutions
of much increased structural strength where the amount of added reinforcing
material is very low.
The employed objectives and their sensitivities are based on the stiffness
and mass matrices and the method can be extended to any other objectives having the same dependency. The presented hybrid evaluation of the sensitivities
allows optimizing complex models without the need of implementing all applied element types. For many problems, the method is very efficient and large
FEM models can be processed. Even though, computer memory requirements
can become high if sensitivities depend on displacement derivatives which may
limit the model size. The method has been shown to find the optima of convex
design problems with existing analytical solutions where small deviations can
be attributed to the discretization of the numerical models. It does not guarantee finding the global optimum of a given multimodal design problem and the
solution process is thus understood as an automated design process. However,
the process is efficient and robust, the requirements on computer infrastructure and licenses are low, and the improvements over initial designs can be
very high. The improved design solutions are feasible for manufacturing as
final layer thickness, the stacking sequence and the orientation angles are predefined in the parameterization and the resulting reinforcements tend to be
coherent. Isolated reinforcements covering only small areas can usually be
disregarded without significantly changing the structural properties.
The need for realizing the locally reinforced structures designed by the presented method provides a new set of manufacturing related problems to be
solved. The sequential curing process, here applied for the vibrating panel
and the notched plate specimens, has given very good laminate quality and
allows an exact realization of the reinforcements. The experimental results
of the vibrating panel demonstrate high agreement with the simulation which

9.1 Discussion

197

emphasize the quality of the process. Even if it is employed only for a small
number of parts, it potentially can be enhanced to an automated process for
larger quantities.
As stated initially, the search for best structural performance must include
an adaption of the local laminate properties to the local load distributions and
directions, respectively. In this context, the application of quasi-isotropic laminates is suboptimal. Nonetheless, they have prevalently been used as basic
laminates for the presented examples. Starting from a quasi-isotropic laminate, the structural performance improvement by inducing a local anisotropy
with local reinforcements becomes obvious. Figure 9.1 illustrates the local
anisotropy for the vibrating panel. The blue lines represent the material prin-

Figure 9.1: Material principal axes of the vibrating panel


cipal directions of the homogenized laminate for each element. The line length
expresses the ratio of the stiffness in principal direction to the stiffness perpendicular to it wherefore it quantifies the anisotropy. Elements containing points
still have a quasi-isotropic laminate. The plot reminds of the examples with
local varying orientation angles presented in the introduction (see Figure 1.7).
It demonstrates that the application of locally reinforced laminates is feasible
to take advantage of the anisotropic material properties. While some existing
approaches leave one with the additional inverse problem of finding a laminate

198

Discussion and Conclusion

reproducing the found anisotropy distribution, the solution here contains the
specific layup. From the manufacturing point of view, solutions with local reinforcements are easier to realize than solutions with varying orientation angles.

9.2 Outlook
This section contains potential ideas for further research based on the deliverables of this thesis. The advanced tasks are separated into improvement of
the method, extension of the method and enhancement of the manufacturing
process.
Improvement of the Method
The presented method could be more efficient by implementing it in a basic program language. MATLAB is an engineering software which is easy to
learn without fundamental programming knowledge. However, there are some
penalties concerning efficiency and memory allocation. A programming of the
method in a low-level language would enable the optimization of larger models
in less time. However, this cannot be understood as research work since the
basic idea is not different.
Alternatively, further investigations could be done in the field of the applied optimization algorithms. Different strategies for the generation of the
reinforcements can be defined. Other algorithms may be faster and the quality
of the obtained solutions could be better. An implementation of the method of
feasible direction [192, 173] would allow considering side constraints.
Extension of the Method
Here, eigenfrequency, compliance, buckling, strength and combinations of
them have been considered. More objective functions can be formulated as
long as they depend on the stiffness- and mass matrix derivatives with respect
to the thicknesses. An interesting task could be to additionally implement
the thermo-mechanical behavior of the material. CFRP-materials have low or
even negative expansion coefficients in fiber direction wherefore they are used
for structures with thermal-dimensional stability. The anisotropic properties
of the materials can be used to specifically design the thermo-mechanical behavior of the structure.
With the increasing application of laminated composites, design methods
considering fatigue requirements become more important. Local reinforcements can basically be helpful to improve the fatigue behavior and the implementation of an adequate objective function could be interesting.

9.2 Outlook

199

A big concern regarding the usage of laminated structure is the afterimpact behavior. Local damages are not as easy to repair as for example for
metallic structures. Furthermore, the resulting strength and stiffness is hard
to be appraised since the damage is not necessarily visible nor can it be modeled. A possibility to restore the original properties, or at least to make them
determinable, is to eliminate the damage and recover them with local reinforcements. With the presented sequential curing process, the reinforcements
can be attached easily. Such repairing techniques are already used today. The
presented method could help to design the repairing patches ideally.
Enhancement of the Manufacturing Process
From the perspective of the author, the field with the greatest potential
for further research is the enhancement of the manufacturing process. As
demonstrated well enough, the presented locally reinforced designs have an
outstanding structural behavior. The proposed manufacturing process which
has mainly be carried out manually is adequate for low piece structures. For
larger units, an automation of the process is needed. Mass production of laminated structures is usually done with Resin Transfer Molding (RTM) processes.
However, PREPREG structures have outstanding laminate properties in terms
of fiber volume content. An accuracy that have been shown in the example
of the vibrating plate can barely be achieved using a RTM process. An automation of the placement of the reinforcements can additionally enhance the
quality. Basically, RTM and PREPREG process can be combined in case that the
matrix resin systems are compatible. The basic structure can be manufactured
with RTM and the reinforcements are made of PREPREG. The idea of individually design the structure to the respective loads which are dependent on the
customer requirements could be interesting for future research.

200

Discussion and Conclusion

Appendix A

Material Properties
A.1 T300/Epoxy
E 11
E 22
G 12
G 23
12

138
9
4.25
3.25
0.34
1556

GPa
GPa
GPa
GPa

k g/m3

Xt
Xc
Yt
Yc
S

2330
1500
47
235
79

MPa
MPa
MPa
MPa
MPa

Table A.1: Material properties for a standard T300/Epoxy laminate (regular)

E 11
E 22
G 12
G 23
12

130
9
4.25
3.25
0.34
1556

GPa
GPa
GPa
GPa

k g/m3

Xt
Xc
Yt
Yc
S

2330
1500
47
235
79

MPa
MPa
MPa
MPa
MPa

Table A.2: Material properties for a standard T300/Epoxy laminate (used for
the notched plate: based on a comparison between simulation and experiments,
the modulus E 11 has been slightly reduced)

ii

Material Properties

A.2 Aluminum 2014-T6


E

73.1
0.33
28
2800

GPa

GPa
k g/m3

R0.2

414

MPa

Table A.3: Material properties for aluminum 2014-T6

Appendix B

Stiffness Matrix
Sensitivities
The stiffness matrix sensitivity of a finite element with respect to the thickness
t l consists of the stiffness sensitivities of the sub-parts, namely the membrane
(m), the coupling (c), the bending (b) and the transverse shear stiffness (s).
dk dkm dk c dkb dks
=
+
+
+
dt l
dt l
dt l
dt l
dt l

(B.1)

The corresponding parts are given by equations B.2 through B.7. The sensitivities of the membrane parts are simply given by

dkm
(B.2)
= BT
m Q l Bm d A
dt l
A

The coupling part is


 


  j 1
!!
j
1


k c l
z0
z0
c
=
kj
t k z0
t k z0

t l
t l
t l
j =1
k =1
k =1






l
l
1
z0
z0
c 
t k z0 1
t k z0

+ kl

t
t l
l
k =1
k =1
 




!!
j
j
1
n


z0
z0
c
kj
t k z0 1
t k z0 1

+
t l
t l
j = l +1
k =1
k =1
with
c

kj =


A


T
BT
m Q j Bb + Bb Q j Bm d A

(B.3)

(B.4)

iv

Stiffness Matrix Sensitivities

The sensitivities of the bending part yield


kb
t l

l
1

j =1
b

+k l
+

b
kj

l

k =1

n

j = l +1



j

k =1

2

t k z0
2

t k z0
b

kj



j

k =1

z0
t l

z0

t l
 

k =1

l
1

2

t k z0

2

z0

z0

!!

t l

t l
2
  j 1
!!

z0
z0
1
t k z0
1
(B.5)

t l
t l
k =1

2

t k z0

  j 1

t k z0

k =1

with
b

kj =


A

BT
b Q j Bb d A

The part arising from the transverse shear stiffness is simply



dks
S
= BT
s Ql B s d A
dt l
A

(B.6)

(B.7)

Appendix C

Beam Finite Elements


The beam finite elements used for the verification in Chapter 5 have been
taken from Cook et al. [28]. Each element has two translational and two ro-

y,v
1

2
x

v1

E,L,

v2

Figure C.1: Beam finite element

tational degrees of freedom. The corresponding sketch is given in Figure C.1.


E denotes the Youngs modulus, I z (x) the area moment of inertia, L the beam
length, the material density, A the beam section area and P the axial load.
For the mass matrix m, a lumped mass formulation has been taken which neglects the rotary inertia. This is feasible if the rotation is small, e.g. for linear
vibrations. The formulation of stiffness-, mass-, and geometric stiffness matrix
is given below.

vi

Beam Finite Elements

Stiffness Matrix

12

2
L
6

EI b
L
k=

12
L

L2

12

L2
6

4
6

L
2

L
12
L2
6
L

(C.1)

Mass Matrix

1
AL
0
m=
2 0
0

0
0
0
0

0
0
1
0

0
0

0
0

(C.2)

Stress Stiffness Matrix

36
P
3L

k =
30L 36
3L

3L
4L2
3L
L2

36
3L
36
3L

3L
2
L
3L

4L

(C.3)

Appendix D

Mode Comparison of the


Vibrating Panel
The eigenmode shapes of the vibrating panel in the basic and the reinforced
configuration presented in Section 8.1 are contrasted here. Left hand side Figures show the results of the FEM-analysis. The experimental results obtained
with the scanning vibrometer are shown on the right side.

D.1 Modes of the Basic Panel


Figures D.1 to D.6 compare the first six eigenmode shapes of the basic panel.

(a) simulation

(b) experiment

Figure D.1: 1st eigenmode of the panel with basic laminate

viii

Mode Comparison of the Vibrating Panel

(a) simulation

(b) experiment

Figure D.2: 2nd eigenmode of the panel with basic laminate

(a) simulation

(b) experiment

Figure D.3: 3rd eigenmode of the panel with basic laminate

(a) simulation

(b) experiment

Figure D.4: 4th eigenmode of the panel with basic laminate

D.1 Modes of the Basic Panel

(a) simulation

ix

(b) experiment

Figure D.5: 5th eigenmode of the panel with basic laminate

(a) simulation

(b) experiment

Figure D.6: 6th eigenmode of the panel with basic laminate

Mode Comparison of the Vibrating Panel

D.2 Modes of the Reinforced Panel


Figures D.7 to D.11 compare the first five eigenmode shapes of the basic panel.
The 6th eigenmode is not presented because it was not found experimentally.

(a) simulation

(b) experiment

Figure D.7: 1st eigenmode of the panel with reinforcements

(a) simulation

(b) experiment

Figure D.8: 2nd eigenmode of the panel with reinforcements

D.2 Modes of the Reinforced Panel

(a) simulation

xi

(b) experiment

Figure D.9: 3rd eigenmode of the panel with reinforcements

(a) simulation

(b) experiment

Figure D.10: 4th eigenmode of the panel with reinforcements

(a) simulation

(b) experiment

Figure D.11: 5th eigenmode of the panel with reinforcements

xii

Mode Comparison of the Vibrating Panel

Acronyms
AE
Acoustic Emission is a test method to identify micro-mechanical effects in continua.
It detects damage events by means of measuring acoustic wave propagation.

179, 186190,
194

Carbon Fiber Reinfoced Plastics is common composite material of a polymer matrix and carbon reinforcement fibers.

2, 3, 9, 63,
171, 174, 178,
188, 198

The Classical Laminate Theory is the


state-of-the art calculus for thin walled
laminated composites.
A fundamental
overview of the theory is given in [6, 72].

1012, 26, 27,


30, 41, 43, 109

Digital Image Correlation is an optical


method to measure changes in images
which is used in engineering to measure
strain or deformation. It determines the
difference between two images taking the
correlation between the gray-scale values
of the pixels.

179, 180, 182


184, 186, 188
190, 194

CFRP

CLT

DIC

xiv

Acronyms

FEM
Finite Element Method is a numerical
technique for finding approximate solution
of physical problems that are based on partial differential equations. Fundamental
and detailed explanations of the FEM focusing on structural analysis can be found
in literature [149, 106, 191, 29, 139]

vii, 2, 20, 35
38, 84, 109,
131, 188, 196

First Ply Failure is the event of first cracking in one ply of the laminate. Damage
predicted by the Tsai-Hill and the Tsai-Wu
failure criterion is called First Ply Failure.

9, 155,
158, 178,
181, 183,
186190,
194, 196

Fiber Reinfoced Plastics is a composite material made of a polymer matrix and reinforcement fibers which are typically of carbon, glass or aramid.

24

Genetic Algorithms are a class of stochastic algorithms that are based on the mechanisms of natural evolution. The algorithm
is discussed in detail by Holland [60].

12, 13, 1517,


99

The Ghost Layer Method is the method for


the generation of local laminate reinforcements presented in this thesis.

89, 99, 105,


115, 118, 121,
123127, 129,
130, 132134,
136, 150, 151,
153, 161, 173,
177, 178, 193,
195

FPF
157,
179,
184,
193,

FRP

GA

GLM

Acronyms

xv

MATLAB
MATLAB is a numerical computing environment which is widely used in academic
and research institutions as well as industrial enterprises [112].

83, 84, 99,


105, 106, 198

NASTRAN is a commercial finite element analysis software based on FORTRAN


which was originally developed for NASA
[117].

16, 17, 41, 83,


84, 105, 106

PREPREG is a type of semi-finished composite where the fibers are already preimpregnated with the matrix system.

4, 161, 162,
167, 174, 199

The Pseudo Strength Function unites the


local strength expressions into a global objective function and allows a differentiation
to obtain strength sensitivities.

7678,
155, 157

Resin Transfer Molding is a process for the


production of laminated composites where
the dry fibers are placed in a mold and the
resin system is injected afterwards.

161, 199

NASTRAN

PREPREG

PSF

RTM

82,

xvi

Acronyms

Bibliography
[1] M. M. Abdalla, S. Setoodeh, and Z. Grdal. Design of variable stiffness
composite panels for maximum fundamental frequency using lamination
parameters. Composite Structures, 81(2):283291, 2007.
[2] S. Abrate. Optimal design of laminated plates and shells. Composite
Structures, 29:269286, 1994.
[3] S. Adali. Design sensitivity analysis of an antisymmetric angle-ply laminate. Engineering Optimization, 7(1):6983, 1983.
[4] S. Adali and K. J. Duffy. Design of antisymmetric hybrid laminates for
maximum buckling load: I. optimal fibre orientation. Composite Structures, 14(1):4960, 1990.
[5] D. B. Adams, L. T. Watson, Z. Grdal, and C. M. Anderson-Cook. Genetic algorithm optimization and blending of composite laminates by locally reducing laminate thickness. Advances in Engineering Software,
35(1):3543, 2004.
[6] J. E. Ashton, J. C. Halpin, and P. H. Petit. Primer on Composite Materials: Analysis. Technomic Press, Westport, CT, 1969.
[7] M. Autio. Determining the real lay-up of a laminate corresponding to
optimal lamination parameters by genetic search. Structural and Multidisciplinary Optimization, 20:301310, 2000.
[8] R.C. Averill and J.N. Reddy. Behaviour of plate elements based on
the first-order shear deformation theory. Engineering Computations,
7(1):5774, 1990.
[9] S. A. Barakat and G. A. Abu-Farsakh. The use of an energy-based criterion to determine optimum configurations of fibrous composites. Composites Science and Technology, 59(12):18911899, 1999.

xviii

BIBLIOGRAPHY

[10] W. Becker, P. P. Jin, and P. Neuser. Interlaminar stresses at the free


corners of a laminate. Composite Structures, 45(2):155162, 1999.
[11] R. Belevicius. Shape optimization of laminated orthotropic plate structures. Mechanics of Composite Materials, 29:406413, 1994.
[12] W. Beluch. Evolutionary identification and optimization of composite
structures. In III European Conference on Computational Mechanics,
pages 446446. Springer Netherlands, 2006.
[13] A. Ben-Tal, M. Kocvara, A. Nemirovski, and J. Zowe. Free material
design via semidefinite programming: The multiload case with contact
conditions. Conditions, Optimization Laboratory, Faculty of Industrial
Engineering and Management, Technion - Israel Institute of Technology,
9:813832, 1997.
[14] M. P. Bendse. Optimal shape design as a material distribution problem.
Structural and Multidisciplinary Optimization, 1(4):193202, 1989.
[15] M. P. Bendse and N. Kikuchi. Generating optimal topologies in structural design using a homogenization method. Computer Methods in Applied Mechanics and Engineering, 71:197224, 1988.
[16] C. W. Bert. Optimal design of a composite-material plate to maximize its
fundamental frequency. Journal of Sound and Vibration, 50(2):229237,
1977.
[17] J. D. Brennan and T. J. Sherwood. Compaction of prepreg plies on composite laminate structures, April 2010.
[18] H. Bruck, S. McNeill, M. Sutton, and W. Peters. Digital image correlation
using newton-raphson method of partial differential correction. Experimental Mechanics, 29:261267, 1989.
[19] R. H. Byrd, J. C. Gilbert, and J. Nocedal. A trust region method based
on interior point techniques for nonlinear programming. Mathematical
Programming, 89:149185, 1996.
[20] R. H. Byrd, M. E. Hribar, and J. Nocedal. An interior point algorithm
for large scale nonlinear programming. SIAM Journal on Optimization,
9:877900, 1997.
[21] K. J. Callahan and G. E. Weeks. Optimum design of composite laminates
using genetic algorithms. Composites Part B: Engineering, 2(3):149160,
1992.

BIBLIOGRAPHY

xix

[22] A. Cauchy. Method generale pour la resolution des systemes dequations


simultanees. Compt. Rend. Acad. Sci., 25:536538, 1847.
[23] F. Cesari, V. Dal Re, G. Minak, and A. Zucchelli. Damage and residual strength of laminated carbonepoxy composite circular plates loaded
at the centre. Composites Part A: Applied Science and Manufacturing,
38(4):11631173, 2007.
[24] R. R. Chang and T. H. Chiang. Theoretical and experimental predictions
of first-ply failure of a laminated composite elevated floor plate. Proceedings of the Institution of Mechanical Engineers, Part E: Journal of
Process Mechanical Engineering, 224(4):233245, 2010.
[25] T. L. Chao. Minimum weight design of stiffened fibre composite cylinders. Technical report, Air Force Materials Laboratory, Wright-Patterson
Air Force Base, Ohio, AFML TR-69-251, 1969.
[26] S. Chen, P. Guan, and B. Shen. Sensitivity analysis of eigenproblems
for symmetrically laminated composite plates. Computers & Structures,
59:431435, 1996.
[27] T. Chu, W. Ranson, and M. Sutton. Applications of digital-imagecorrelation techniques to experimental mechanics. Experimental Mechanics, 25:232244, 1985.
[28] R. D. Cook. Four-node flat shell element: Drilling degrees of freedom,
membrane-bending coupling, warped geometry, and behavior. Computers & Structures, 50(4):549555, 1994.
[29] R. D. Cook, D. S. Malkus, M. E. Plesha, and R. J. Witt. Concepts and
applications of finite element analysis. John Wiley & Sons Inc, 4 edition,
2001.
[30] R. G. Cuntze and A. Freund. The predictive capability of failure mode
concept-based strength criteria for multidirectional laminates. Composites Science and Technology, 64(3-4):343377, 2004.
[31] S. Daggumati, I. De Baere, W. Van Paepegem, J. Degrieck, J. Xu, S.V.
Lomov, and I. Verpoest. Local damage in a 5-harness satin weave composite under static tension: Part I experimental analysis. Composites
Science and Technology, 70(13):19261933, 2010.
[32] S. B. Dong and F. K. W. Tso. On a laminated orthotropic shell theory
including transverse shear deformation. Journal of Applied Mechanics,
39(4):10911097, 1972.

xx

BIBLIOGRAPHY

[33] K. J. Duffy and S. Adali. Design of antisymmetric hybrid laminates for


maximum buckling load: II. optimal layer thickness. Composite Structures, 14(2):113124, 1990.
[34] H. Engels, W. Hansel, and W. Becker. Optimal design of hole reinforcements for composite structures. Mechanics of Composite Material,
38(5):417428, 2002.
[35] B. G. Falzon, G. P. Steven, and Y. M. Xie. Shape optimization of interior
cutouts in composite panels. Structural and Multidisciplinary Optimization, 11:4349, 1996.
[36] J. S. R. Fellow, K. Sheelavant, and B. Bombale. Concept optimal design
of composite fan blades. ASME Turbo Expo 2011 June 9-13, 2011 Vancouver, Canada, 2011.
[37] R. Fletcher and C. M. Reeves. Function minimization by conjugate gradients. The Computer Journal, 7(2):149154, 1964.
[38] J. Foldager, J. S. Hansen, and N. Olhoff. A general approach forcing
convexity of ply angle optimization in composite laminates. Structural
Optimization, 16:201211, 1998.
[39] R. Foye. Advanced design concepts for advanced composite air frames.
Technical Report 1-11, Air Force Materials Laboratory, Wright-Patterson
Air Force Base, Ohio, AFML-TR-68-91, 1968.
[40] H. Fukunaga. On laminate configurations for simultaneous failure.
Journal of Composite Materials, 22(3):271286, 1988.
[41] H. Fukunaga and H. Sekine. Optimum design of composite structures
for shape, layer angle and layer thickness distributions. Journal of Composite Materials, 27:14791492, 1993.
[42] H. Fukunaga, H. Sekine, M. Sato, and A. Iino. Buckling design of symmetrically laminated plates using lamination parameters. Computers &
Structures, 57(4):643649, 1995.
[43] H. Fukunaga and G.N. Vanderplaats. Strength optimization of laminated composites with respect to layer thickness and/or layer orientation
angle. Computers & Structures, 40(6):14291439, 1991.
[44] L. V. Gibiansky and O. Sigmund. Multiphase composites with extremal bulk modulus. Journal of the Mechanics and Physics of Solids,
48(3):461498, 2000.

BIBLIOGRAPHY

xxi

[45] M. Giger, D. Keller, and P. Ermanni. A graph-based parameterization


concept for global laminate optimization. Structural and Multidisciplinary Optimization, 36(3):289305, 2008.
[46] H. Goldstein. Classical Mechanics. Addison-Wesley, 2001.
[47] J. E. Gordon. The new science of strong materials: or, Why you dont fall
through the floor. Pelican books. Penguin, 1991.
[48] B. G. Green, M. R. Wisnom, and S. R. Hallett. An experimental investigation into the tensile strength scaling of notched composites. Composites
Part A: Applied Science and Manufacturing, 38(3):867878, 2007.
[49] J. L. Grenestedt. Layup optimization and sensitivity analysis of the
fundamental eigenfrequency of composite plates. Composite Structures,
12(3):193209, 1989.
[50] J. L. Grenestedt. Layup optimization against buckling of shear panels.
Structural Optimization, 3:115120, 1991.
[51] Z. Grdal, R. Haftka, and P. Hajela. Design and Optimization of Laminated Composite Materials. John, 1999.
[52] R. T. Haftka and Z. Grdal. Elements of structural optimization. Solid
mechanics and its applications. Kluwer Academic Publishers, 1992.
[53] W. Hansel and W. Becker. Layerwise adaptive topology optimization of
laminate structures. Engineering Computations, 16(7):841851, 1999.
[54] W. Hansel, A. Treptow, W. Becker, and B. Freisleben. A heuristic and
genetic topology optimization algorithm for weight-minimal laminate
structures. Composite Structures, 58(2):287294, 2002.
[55] C. E. Harris and D. H. Morris. Role of delamination and damage development on the strength of thick notched laminates. In W.S Johnson,
editor, Delamination and debonding of materials, pages 424447. ASTM
STP 876, Philadelphia, 1985.
[56] Z. Hashin. Failure criteria for unidirectional fiber composites. Journal
of Applied Mechanics, 47(2):329334, 1980.
[57] R. Hill. A theory of the yielding and plastic flow of anisotropic metals.
Proceedings of the Royal Society of London. Series A. Mathematical and
Physical Sciences, 193(1033):281297, 1948.

xxii

BIBLIOGRAPHY

[58] M. J. Hinton, A. S. Kaddour, and P. D. Soden. Failure criteria in fibre


reinforced polymer composites: the world-wide failure exercise. Elsevier,
2004.
[59] O. Hoffman. The brittle strength of orthotropic materials. Journal of
Composite Materials, 1(2):200206, 1967.
[60] J. H. Holland. Adaptation in natural and artificial systems. MIT Press,
Cambridge, MA, USA, 1975.
[61] H. R. E. M. Hrnlein, M. Kocvara, and R. Werner. Moped: An integrated
designer tool for material optimization. Technical Report 279, Universitt Erlangen, Institut fr angewandte Mathematik, 2001.
[62] J. Huang and R. T. Haftka. Optimization of fiber orientations near a hole
for increased load-carrying capacity of composite laminates. Structural
and Multidisciplinary Optimization, 30(5):335341, 2005.
[63] T. J. R. Hughes, M. Cohen, and M. Haroun. Reduced and selective integration techniques in the finite element analysis of plates. Nuclear
Engineering and Design, 46(1):203222, 1978.
[64] S. T. IJsselmuiden, M. M. Abdalla, and Z. Grdal. Optimization of
variable-stiffness panels for maximum buckling load using lamination
parameters. AIAA JOURNAL, 48(1):134143, 2010.
[65] S. T. IJsselmuiden, M. M. Abdalla, V. K. Pilaka, and Z. Grdal. Design
of variable stiffness composite structures for advanced fibre placement
technology. In SAMPE 2010 - Seattle, WA May 17-20, 2010.
[66] M. Iura and S. N. Atluri. Formulation of a membrane finite element
with drilling degrees of freedom. Computational Mechanics, 9(6):417
428, 1992.
[67] L. Iuspa and E. Ruocco. Optimum topological design of simply supported
composite stiffened panels via genetic algorithms. Computers & Structures, 86(1718):17181737, 2008.
[68] D. Ivanov, S. Ivanov, S. Lomov, and I. Verpoest. Strain mapping analysis
of textile composites. Optics and Lasers in Engineering, 47(3-4):360
370, 2009.
[69] D. C. Jegley, B. F. Tatting, and Z. Grdal. Optimization of elastically
tailored tow-placed plates with holes. In 44th AIAA/ASME/ASCE/AHS

BIBLIOGRAPHY

xxiii

Structures, Structural Dynamics, and Materials Conference, American


Institute of Aeronautics and Astronautics, Norfolk, VA., 2003.
[70] L. Johansen and E. Lund. Optimization of laminated composite structures using delamination criteria and hierarchical models. Structural
and Multidisciplinary Optimization, 38:357375, 2009.
[71] L. Johansen, E. Lund, and J. Kleist. Failure optimization of geometrically linear/nonlinear laminated composite structures using a two-step
hierarchical model adaptivity. Computer Methods in Applied Mechanics
and Engineering, 198(3032):24212438, 2009.
[72] R. M. Jones. Mechanics of Composite Materials. Hemisphere Publishing
Corporation, New York, 1975.
[73] T. Y. Kam and M. D. Lai. Multilevel optimal design of laminated composite plate structures. Computers & Structures, 31(2):197202, 1989.
[74] T.Y. Kam and F.M. Lai. Experimental and theoretical predictions of firstply failure strength of laminated composite plates. International Journal
of Solids and Structures, 36(16):23792395, 1999.
[75] B. L. Karihaloo and F. I. Niordson. Optimum design of vibrating cantilevers. Journal of Optimization Theory and Applications, 11:638654,
1973.
[76] B. L. Karihaloo and R. D. Parbery. Optimal design of multi-purpose
beam-columns. Journal of Optimization Theory and Applications,
27:439448, 1979.
[77] B. L. Karihaloo and R. D. Parbery. Optimal design of beam-columns
subjected to concentrated moments. Engineering Optimization, 5(1):59
65, 1980.
[78] D. Keller. Evolutionary Design of Laminated Composite Structures. In
Diss. ETH. No. 19011, ETH Zurich, Switzerland, 2010.
[79] D. Keller.
Global laminate optimization on geometrically partitioned shell structures. Structural and Multidisciplinary Optimization,
43(3):353368, 2011.
[80] A. Khani, S. T. IJsselmuiden, M. M. Abdalla, and Z. Grdal. Design
of variable stiffness panels for maximum strength using lamination parameters. Composites Part B: Engineering, 42(3):546552, 2011.

xxiv

BIBLIOGRAPHY

[81] P. Khosravi and R. Sedaghati. Design of laminated composite structures


for optimum fiber direction and layer thickness, using optimality criteria. Structural and Multidisciplinary Optimization, 36:159167, 2008.
[82] N. S. Khot, V. B. Venkayya, C. D. Johnson, and V. A. Tischler. Optimization of fiber reinforced composite structures. International Journal of
Solids and Structures, 9(10):12251236, 1973.
[83] F. Kikuchi and K. Ishii. An improved 4-node quadrilateral plate bending
element of the reissner-mindlin type. Computational Mechanics, 23:240
249, 1999.
[84] J.-S. Kim. Development of a user-friendly expert system for composite
laminate design. Composite Structures, 79(1):7683, 2007.
[85] N. Kogiso, L. T. Watson, Z. Grdal, and R. T. Haftka. Genetic algorithms
with local improvement for composite laminate design. Structural and
Multidisciplinary Optimization, 7:207218, 1994.
[86] M. Kocvara and M. Stingl. Free material optimization for stress constraints. Structural and Multidisciplinary Optimization, 33:323335,
2007.
[87] G. Kress. Shape optimization of a flywheel. Structural and Multidisciplinary Optimization, 19:7481, 2000.
[88] G. Kress, P. Naeff, M. Niedermeier, and P. Ermanni. Onsert strength
design. International Journal of Adhesion and Adhesives, 24(3):201209,
2004.
[89] B. P. Kristinsdottir, Z. B. Zabinsky, M. E. Tuttle, and S. Neogi. Optimal
design of large composite panels with varying loads. Composite Structures, 51(1):93102, 2001.
[90] M. Laitinen, H. Lahtinen, and S.-G. Sjlind. Transverse shear correction
factors for laminates in cylindrical bending. Communications in Numerical Methods in Engineering, 11(1):4147, 1995.
[91] R. Le Riche and R.T. Haftka. Improved genetic algorithm for minimum
thickness composite laminate design. Composites Engineering, 5(2):143
161, 1995.
[92] B. Liu, R. T. Haftka, and P. Trompette. Maximization of buckling loads of
composite panels using flexural lamination parameters. Structural and
Multidisciplinary Optimization, 26:2836, 2004.

BIBLIOGRAPHY

xxv

[93] B. Liu, R.T. Haftka, and M.A. Akgn. Two-level composite wing structural optimization using response surfaces. Structural and Multidisciplinary Optimization, 20:8796, 2000.
[94] K.-S. Liu and S. W. Tsai. A progressive quadratic failure criterion for a
laminate. Composites Science and Technology, 58(7):10231032, 1998.
[95] Y. Liu, F. Jin, and Q. Li. A strength-based multiple cutout optimization
in composite plates using fixed grid finite element method. Composite
Structures, 73(4):403412, 2006.
[96] S. V. Lomov, D. S. Ivanov, I. Verpoest, M. Zako, T. Kurashiki, H. Nakai,
J. Molimard, and A. Vautrin. Full-field strain measurements for validation of meso-fe analysis of textile composites. Composites Part A: Applied
Science and Manufacturing, 39(8):12181231, 2008.
[97] C. S. Lopes, Z. Grdal, and P. P. Camanho. Variable-stiffness composite
panels: Buckling and first-ply failure improvements over straight-fibre
laminates. Computers & Structures, 86(9):897907, 2008.
[98] C. S. Lopes, Z. Grdal, and P. P. Camanho. Tailoring for strength of
composite steered-fibre panels with cutouts. Composites Part A: Applied
Science and Manufacturing, 41(12):17601767, 2010.
[99] A. E. H. Love. The small free vibrations and deformation of a thin elastic shell. Philosophical Transactions of the Royal Society of London. A,
179:pp. 491546, 1888.
[100] E. Lund. Finite Element Based Design Sensitivity Analysis and Optimization. PhD thesis, Aalborg University, Denmark, 1994.
[101] E. Lund. Buckling topology optimization of laminated multi-material
composite shell structures. Composite Structures, 91(2):158167, 2009.
[102] E. Lund and J. Stegmann. On structural optimization of composite shell
structures using a discrete constitutive parametrization. Wind Energy,
8(1):109124, 2005.
[103] E. Lund and J. Stegmann. Eigenfrequency and buckling optimization of
laminated composite shell structures using discrete material optimization. Solid Mechanics and Its Applications, 3:147156, 2006.
[104] R. H. MacNeal. A simple quadrilateral shell element. Computers &
Structures, 8(2):175183, 1978.

xxvi

BIBLIOGRAPHY

[105] R. H. MacNeal. Derivation of element stiffness matrices by assumed


strain distributions. Nuclear Engineering and Design, 70(1):312, 1982.
[106] R. H. MacNeal. Finite elements: their design and performance. Marcel
Dekker, INC., 1994.
[107] R. H. MacNeal and R. L. Harder. A refined four-noded membrane element with rotational degrees of freedom. Computers & Structures,
28(1):7584, 1988.
[108] H. C. Mateus, H. C. Rodrlgues, C. M. Mota Soares, and C. A. Mota Soares.
Sensitivity analysis and optimization of thin laminated structures with a
nonsmooth eigenvalue based criterion. Structural Optimization, 14:219
224, 1997.
[109] H. C. Mateus, C. M. Mota Soare, and C. A. Mota Soarest. Buckling
and sensitivity analysis and optimal design of thin laminated structures.
Computers & Structures, 64(1-4):461472, 1997.
[110] H. C. Mateus, C. M. Mota Soares, and C. A. Mota Soares. Sensitivity analysis and optimal design of thin laminated composite structures.
Computers & Structures, 41:501508, 1991.
[111] J.-D. Mathias, X. Balandraud, and M. Grediac. Applying a genetic algorithm to the optimization of composite patches. Computers & Structures,
84:823834, 2006.
[112] The MathWorks, Inc., Natick, Massachusetts, USA. MATLAB 7.12.0,
R2011a edition, 2011.
[113] A. G. M. Michell. The limits of economy of material in frame-structures.
Philosophical Magazine, 8(47):589597, 1904.
[114] C. Mittelstedt and W. Becker. Interlaminar stress concentrations in layered structures: Part I - A selective literature survey on the free-edge
effect since 1967. Journal of Composite Materials, 38(12):10371062,
2004.
[115] J. S. Moita, J. Infante Barbosa, C. M. Mota Soares, and C. A. Mota Soares.
Sensitivity analysis and optimal design of geometrically non-linear laminated plates and shells. Computers and Structures, 76:407420, 2000.
[116] MSC.Software Corporation, 2 Mac Arthur Place, Santa Ana, CA, USA.
MSC.Nastran 2004 Reference Manual, 2003.

BIBLIOGRAPHY

xxvii

[117] MSC.Software Corporation, 2 Mac Arthur Place, Santa Ana, CA, USA.
MSC.Nastran 2008 Reference Manual, 2003.
[118] G. N. Naik, S. GGopalakrishnan, and R. Ganguli. Design optimization of
composites using genetic algorithms and failure mechanism based failure criterion. Composite Structures, 83(4):354367, 2008.
[119] F. I. Niordson. On the optimal design of a vibrating beam. The Quarterly
of Applied Mathematics, 23:47, 1965.
[120] Library of Congress. www.loc.gov, 2011.
[121] N. Olhoff. Maximizing higher order eigenfrequencies of beams with
constraints on the design geometry. Journal of Structural Mechanics,
5(2):107134, 1977.
[122] D.H. Pahr and F.G Rammerstorfer. Experimental and numerical investigations of perforated cfr woven fabric laminates. Composites Science
and Technology, 64(9):14031410, 2004.
[123] B. Pan, K. Qian, H. Xie, and A. Asundi. Two-dimensional digital image
correlation for in-plane displacement and strain measurement: a review.
Measurement Science and Technology, 20(6):062001, 2009.
[124] C. H. Park, W. I. Lee, W. S. Han, and A. Vautrin. Improved genetic algorithm for multidisciplinary optimization of composite laminates. Computers & Structures, 86(1920):18941903, 2008.
[125] J. H. Park, J. H. Hwang, C. S. Lee, and W. Hwang. Stacking sequence
design of composite laminates for maximum strength using genetic algorithms. Composite Structures, 52(2):217231, 2001.
[126] L. Parnas, S. Oral, and . Ceyhan. Optimum design of composite structures with curved fiber courses. Composites Science and Technology,
63(7):10711082, 2003.
[127] N. Pedersen. On design of fiber-nets and orientation for eigenfrequency
optimization of plates. Computational Mechanics, 39:113, 2005.
[128] N. Pedersen. On simultaneous shape and orientational design for eigenfrequency optimization. Structural and Multidisciplinary Optimization,
33:387399, 2007.
[129] P. Pedersen. On sensitivity analysis and optimal design of specially orthotropic laminates. Engineering Optimization, 11(3-4):305316, 1987.

xxviii

BIBLIOGRAPHY

[130] P. Pedersen. On thickness and orientational design with orthotropic


materials. Structural and Multidisciplinary Optimization, 3(2):6978,
1991.
[131] P. Pedersen. Some general optimal design results using anisotropic,
power law nonlinear elasticity. Structural and Multidisciplinary Optimization, 15:7380, 1998.
[132] J. L. Pelletier and S. S. Vel. Multi-objective optimization of fiber reinforced composite laminates for strength, stiffness and minimal mass.
Computers & Structures, 84(29-30):20652080, 2006.
[133] W. Prager. Optimal design of statically determinate beams for given
deflection. International Journal of Mechanical Sciences, 13(10):893
895, 1971.
[134] W. Prager and J. E. Taylor. Problems of optimal structural design. Journal of Applied Mechanics, 35:102106, 1968.
[135] A. Puck and H. Schrmann. Failure analysis of frp laminates by means
of physically based phenomenological models. Composites Science and
Technology, 58(7):10451067, 1998.
[136] A. Rama Mohan Rao and N. Arvind. A scatter search algorithm for stacking sequence optimisation of laminate composites. Composite Structures,
70(4):383402, 2005.
[137] S. S. Rao and K. Singh. Optimum design of laminates with natural
frequency constraints. Journal of Sound and Vibration, 67(1):101112,
1979.
[138] J. N. Reddy. A penalty plate-bending element for the analysis of laminated anisotropic composite plates. International Journal for Numerical
Methods in Engineering, 15(8):11871206, 1980.
[139] J. N. Reddy. An introduction to the finite element method. McGraw-Hill
series in mechanical engineering. McGraw-Hill Higher Education, 2006.
[140] B. Rentsch. Fabrication and measurement of a locally-reinforced composite plate with the scanning vibrometer. Semesters thesis, Centre of
Structure Technologies, ETH Zurich, 2012.
[141] R. C. Reuter.
Concise property transformation relations for an
anisotropic lamina. Journal of Composite Materials, 5(2):270272, 1971.

BIBLIOGRAPHY

xxix

[142] G. I. N. Rozvany and W. Prager. Structural design via optimality criteria:


the Prager approach to structural optimization. Mechanics of elastic and
inelastic solids. Kluwer Academic Publishers, 1989.
[143] B. Schlpfer. Optimierung von Faserverbundstrukturen fr die Raumfahrt. Masters thesis, ETH Zurich, 2008.
[144] B. Schlpfer and G. Kress. A new methodology for the placement of
reinforcement doublers on composite space structures. In ECCM 14 :
14th European Conference on Composite Materials : June 7-10, 2010,
Budapest, Hungary, 2010.
[145] B. Schlpfer and G. Kress. A sensitivity-based parameterization concept for the automated design and placement of reinforcement doublers.
Composite Structures, 94(3):896903, 2012.
[146] L. A. Schmit and B. Farshi. Optimum laminate design for strength and
stiffness. International Journal for Numerical Methods in Engineering,
7:519536, 1973.
[147] L. A. Schmit and B. Farshi. Optimum design of laminated fibre composite
plates. International Journal for Numerical Methods in Engineering,
11:623640, 1977.
[148] B. Schneuwly. Fertigung und Test einer gelochten Platte mit lokalen
Laminatverstrkungen. Semesters thesis, Centre of Structure Technologies, ETH Zurich, 2011.
[149] H.R. Schwarz. Methode der finiten Elemente: eine Einfhrung unter
besonderer Bercksichtigung der Rechenpraxis. Leitfden der angewandten Mathematik und Mechanik. Teubner, 1991.
[150] O. Seresta, Z. Grdal, D. B. Adams, and L. T. Watson. Optimal design of
composite wing structures with blended laminates. Composites Part B:
Engineering, 38(4):469480, 2007.
[151] S. Setoodeh, M. M. Abdalla, S. T. IJsselmuiden, and Z. Grdal. Design
of variable-stiffness composite panels for maximum buckling load. Composite Structures, 87(1):109117, 2009.
[152] A.P. Seyraniant, E. Lund, and N. Olhoff. Multiple eigenvalues in structural optimization problems. Structural Optimization, 8:207227, 1994.

xxx

BIBLIOGRAPHY

[153] O. Sigmund and S. Torquato. Design of materials with extreme thermal


expansion using a three-phase topology optimization method. Journal of
the Mechanics and Physics of Solids, 45(6):10371067, 1997.
[154] S. Sihn, R. Y. Kim, K. Kawabe, and S. W. Tsai. Experimental studies
of thin-ply laminated composites. Composites Science and Technology,
67(6):9961008, 2007.
[155] K. Sivakumar, N.G.R. Iyengar, and Kalyanmoy Deb. Optimum design
of laminated composite plates with cutouts using a genetic algorithm.
Composite Structures, 42(3):265279, 1998.
[156] M.-G. Song, J.-H. Kweon, J.-H. Choi, J.-H. Byun, M.-H. Song, S.-J. Shin,
and T.-J. Lee. Effect of manufacturing methods on the shear strength of
composite single-lap bonded joints. Composite Structures, 92(9):2194
2202, 2010. Fifteenth International Conference on Composite Structures.
[157] G. Soremekun, Z. Grdal, R. T. Haftka, and L. T. Watson. Composite
laminate design optimization by genetic algorithm with generalized elitist selection. Computers & Structures, 79(2):131143, 2001.
[158] G. Soremekun, Z. Grdal, C. Kassapoglou, and D. Toni. Stacking sequence blending of multiple composite laminates using genetic algorithms. Composite Structures, 56(1):5362, 2002.
[159] J. C. Spall. Introduction to stochastic search and optimization : estimation, simulation, and control. Wiley-Interscience series in discrete
mathematics and optimization, 2003.
[160] J. Stegmann. Analysis and Optimization of Laminated Composite Shell
Structures. PhD thesis, Aalborg University, Denmark, 2005.
[161] J. Stegmann and E. Lund. Discrete material optimization of general
composite shell structures. International Journal for Numerical Methods
in Engineering, 62(14):20092027, 2005.
[162] T. R. Tauchert and S. Adibhatla. Design of laminated plates for maximum bending strength. Engineering Optimization, 8(4):253263, 1985.
[163] A. Tessler and J. R. Hughes. An improved treatment of transverse shear
in the mindlin-type four-node quadrilateral element. Computer Methods
in Applied Mechanics and Engineering, 39:311335, 1983.

BIBLIOGRAPHY

xxxi

[164] A. Todoroki and R. T. Haftka. Stacking sequence optimization by a genetic algorithm with a new recessive gene like repair strategy. Composites Part B: Engineering, 29(3):277285, 1998.
[165] A. Todoroki and T. Ishikawa. Design of experiments for stacking sequence optimizations with genetic algorithm using response surface approximation. Composite Structures, 64(34):349357, 2004.
[166] S. W. Tsai. Fundamental Aspects of Fiber Reinforced Plastic Composites.
Wiley Interscience, New York., 1968.
[167] S. W. Tsai. Theory of Composites Design. Think Composites, Dayton,
Ohio, USA, 2008.
[168] S. W Tsai and N. J. Pagano. Invariant properties of composite materials.
Technical report, Air Force Materials Laboratory, Wright-Patterson Air
Force Base, Ohio, AFB, 1968.
[169] S. W. Tsai and E. M. Wu. A general theory of strength for anisotropic
materials. Journal of Composite Materials, 5(1):5880, 1971.
[170] J. M. J. F. van Campen. Optimum lay-up design of variable stiffness
composite structures. PhD thesis, TU Delft, Netherlands, 2011.
[171] J. M. J. F. van Campen, C. Kassapoglou, and Z. Grdal. Generating
realistic laminate fiber angle distributions for optimal variable stiffness
laminates. Composites Part B: Engineering, 43(2):354360, 2012.
[172] F. van Keulen, R.T. Haftka, and N.H. Kim. Review of options for structural design sensitivity analysis. part 1: Linear systems. Computer
Methods in Applied Mechanics and Engineering, 194:32133243, 2005.
[173] G. N. Vanderplaats and F. Moses. Structural optimization by methods of
feasible directions. Computers & Structures, 3(4):739755, 1973.
[174] S. Vlachoutsis. Shear correction factors for plates and shells. International Journal for Numerical Methods in Engineering, 33:15371552,
1992.
[175] M. Walker and R. E. Smith. A technique for the multiobjective optimisation of laminated composite structures using genetic algorithms and
finite element analysis. Composite Structures, 62(1):123128, 2003.
[176] R. A. Waltz, J. L. Morales, J. Nocedal, and D. Orban. An interior algorithm for nonlinear optimization that combines line search and trust
region steps. Mathematical Programming, 107:391408, 2006.

xxxii

BIBLIOGRAPHY

[177] R. I. Watkins and A. J. Morris. A multicriteria objective function optimization scheme for laminated composites for use in multilevel structural optimization schemes. Computer Methods in Applied Mechanics
and Engineering, 60(2):233251, 1987.
[178] J. M. Whitney. Shear correction factors for orthotropic laminates under
static load. Journal of Applied Mechanics, 40(1):302304, 1973.
[179] H. M. Wigger and W. Becker. Inplane stress singularities at the interface
corner of a bimaterial junction. Composite Structures, 69(2):193 199,
2005.
[180] H. M. Wigger and W. Becker. Characterization of inplane loaded
anisotropic interface corners with the boundary finite element method.
Computational Mechanics, 37:153162, 2006.
[181] H. M. Wigger and W. Becker. Investigation of anisotropic reinforcment
patch corners. Composites, 38:11311140, 2007.
[182] H. M. Wigger and W. Becker. Closed-form analysis of the stress field at
rectangular corners of reinforcement patches. Mechanics of Advanced
Materials and Structures, 15(2):104116, 2008.
[183] E. Winkler and H. Steger. Einsatz der Scanning-Laservibrometrie zur
Messung von Schallschnelleverteilungen an Maschinen und Aggregaten.
Technical report, Polytec GmbH, Waldbronn, 2004.
[184] M. R. Wisnom, B. Green, W.-G. Jiang, and S. R. Hallett. Specimen size
effects on the notched strength of composite laminates loaded in tension.
In 16th International Conference on Composite Materials, 2007.
[185] W. H. Wittrick. Rates of change of eigenfvalues, with reference to buckling and vibration problems. Journal of the Royal Aeronautical Society,
66:590591, 1962.
[186] N. Zehnder. Global Optimization of Laminated Structures. In Diss. ETH.
No. 17573, ETH Zurich, Switzerland, 2008.
[187] N. Zehnder and P. Ermanni. A methodology for the global optimization
of laminated composite structures. Composite Structures, 72(3):311320,
2006.
[188] N. Zehnder and P. Ermanni. Optimizing the shape and placement of
patches of reinforcement fibers. Composite Structures, 77(1):19, 2007.

BIBLIOGRAPHY

xxxiii

[189] M. Zhou, R. Fleury, and M. Kemp. Optimization of composite recent


advances and application. In Proc. 13th AIAA/ISSMO Multidisciplinary
Analysis and Optimization Conference, Fort Worth, Texas, 2011.
[190] O. C. Zienkiewicz, R. L. Taylor, and J. M. Too. Reduced integration technique in general analysis of plates and shells. International Journal for
Numerical Methods in Engineering, 3(2):275290, 1971.
[191] O.C. Zienkiewicz and R.L. Taylor. The Finite Element Method: Solid
mechanics. Number Bd. 2 in The Finite Element Method. ButterworthHeinemann, 2000.
[192] G. Zoutendijk. Methods of Feasible Direction. Elsevier, 1960.
[193] J. Zowe, M. Kocvara, and M. P. Bendse. Free material distribution vie
mathematical programming. In T.M. Liebling and D. de Werra, editors,
Mathematical Programming, volume 79 of B, pages 445466. Elsevier
Science B.V., 1997.

xxxiv

BIBLIOGRAPHY

Own publications
[1] B. Schlpfer and G. Kress. A new methodology for the placement of reinforcement doublers on composite space structures, in Proceedings of the
14th European Conference on Composite Materials (ECCM14), Budapest,
Hungary, 2010.
[2] B. Schlpfer and G. Kress. A pseudo strength function for the generation of local laminate reinforcement doublers, in Proceedings of the 18th
International Conference on Composite Materials (ICCM18), Jeju, Korea,
2011.
[3] B. Schlpfer and G. Kress. A sensitivity-based parameterization concept for the automated design and placement of reinforcement doublers,
Composite Structures, 94(3): 896903, 2012.
[4] B. Schlpfer, B. Rentsch and G. Kress. Specific design of laminated composites regarding dynamic behavior by the application of local reinforcements, Composite Structures, 99: 433442, 2013.
[5] B. Schlpfer and G. Kress. Optimal design and testing of laminated
light-weight composite structures with local reinforcements considering
strength constraints. Part I: Design, submitted to Composites Part A:
Applied Science and Manufacturing.
[6] B. Schlpfer, A.J. Brunner and G. Kress. Optimal design and testing of
laminated light-weight composite structures with local reinforcements
considering strength constraints. Part II: Testing, submitted to Composites Part A: Applied Science and Manufacturing.
[7] B. Schlpfer and G. Kress. Advanced design of laminate composites with
local anisotropies by employing local laminate reinforcements, submitted
to Computational Methods in Applied Engineering.

xxxvi

OWN PUBLICATIONS

Curriculum Vitae

P ERSONAL
Name :
Date of Birth :
Nationality :
Contact :

Benjamin Georg Schlpfer


September 13, 1983
Swiss
benjamin.schlaepfer@alumni.ethz.ch

E DUCATION
Ph.D., Institute for Mechanical Systems, ETH Zrich

12/2008 12/2012

Master Study in Mechanical Engineering, ETH Zrich


Focus on Aerospace Engineering and Lightweight Structures

03/2007 09/2008

Bachelor Study in Mechanical Engineering, ETH Zrich


Focus on Structure Mechanics

10/2003 07/2006

Gymnasium, Kantonsschule Trogen


Focus on Mathematics & Physics

08/1998 05/2002

W ORKING E XPERIENCE
IT-Administrator, Centre of Structure Technologies, ETH Zrich

12/2008 12/2012

Internship, Oerlikon Components AG Zrich


Preliminary Design of Satellite Structures

08/2006 02/2007

For all those who came until the end, I created a flipbook at the bottom of the
left hand side pages. You will see growing the 90 reinforcement layer of the
panel discussed in Section 6.1. Take your thumb and flip backwards through
the pages. Enjoy!

You might also like