You are on page 1of 225

Predicting Residual Stresses due

to Solidi cation in Cast Plastic


Plates

Vlado Tropsa

Thesis submitted for the


Degree of Doctor of Philosophy of the University of London
and
Diploma of Imperial College

Imperial College
OF SCIENCE TECHNOLOGY AND MEDICINE

March 2001
Abstract
Thermal processing is an important stage in manufacturing of polymeric products
which may considerably a ect their behaviour. During the cooling from an elevated
temperature to room temperature, strains become \frozen-in" the material. These
frozen-in strains lead to residual stresses and distortions of the nal product.
A method is proposed for the quantitative prediction of the residual stresses in
polymeric materials. A \residual" temperature eld is introduced to describe the
relationship between thermal history and frozen-in strains. This temperature eld,
when applied as an actual temperature distribution, produces thermal stresses and
distortions equal to those caused by residual stresses. The numerical algorithm
for determining the residual temperature eld is described in detail and the Finite
volume Method is used to carry out the thermo-elastic analysis.
Distortions due to residual stresses are obvious in the production of cast plastic
plates. In a typical casting process, the material is rst polymerised at a high tem-
perature and then cooled to ambient temperature. After cooling, the plate is dis-
torted to an extent dependent on the degree of asymmetrical cooling. In a perfectly
symmetrical case no macroscopic distortion is evident but the plate is internally
strained. If the internal equilibrium of such plate is disturbed by sectioning, then
macroscopic distortion will occur. This distortion is proportional to the magnitude
of stresses in the removed section.
The frozen-in strain model is validated through a series of experiments in which
lled PMMA plates are cooled in a controlled environment. During the tests, tem-
peratures are recorded at a number of locations on the plate surfaces. The measured
temperature histories are then used in the numerical analysis to obtain the residual
temperature eld. Comparisons between experimental and numerically predicted
distorted shapes show very good agreement.
2
Acknowledgments
I would like to express my gratitude to my supervisors Dr. Alojz Ivankovic and
Professor Gordon Williams for their guidance and advice during this research, and
even more for the patience and encouragement that I have been constantly receiving
from them.
Special thanks must go to my good friend and colleague Dr. Hrvoje Jasak for his
prompt and valuable help on many occasions. I wish to thank my current supervisor
Professor Tony Kinloch for being very considerate in the past few months, and my
Professors from the University of Zagreb, Branko Matijasevic and Nikola Serman,
for their continuous support.
I am grateful to my good friends, many of them past and present members of the
Strength of Materials Section, particularly Iain Ainsworth, Paul Davis, Mark Dou-
glas, Ioannis Georgiou, Jan Graham, Chris Greenshields, Stuart Hillmansen, Josip
Jagust, Aleksandar Karac, Pat Leevers, Matthew Little, Kuntinee Maneeratana,
Steve Ritchie and Greg Venizelos, for their helpful discussions and contributions to
this work. I would like to extend my gratitude to the administration sta , especially
Maggie Dean, Claire Ferguson, Vernette Rice for their help on many aspects, and
also to members of the technical sta for their assistance.
The nancial support provided by DuPont Company is gratefully acknowledged.
I am dedicating this work to my parents Helana and Stjepan, and dear sister Vesna
to whom I cannot thank enough.

3
Contents

Abstract 2

Acknowledgments 3

Contents 4

List of gures 9

List of tables 14

Abbreviations 15

Nomenclature 16

1 Introduction 19
1.1 General Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . 19
1.2 Computational Methods . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.3 Basic Principles of Linear Viscoelasticity . . . . . . . . . . . . . . . . 28
1.4 Previous and Related Studies . . . . . . . . . . . . . . . . . . . . . . 32
1.5 Layout of the Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
2 Viscoelastic Stress Analysis in a Flat Plate 42
2.1 Thermo-Elastic Analysis of a Flat Plate . . . . . . . . . . . . . . . . 42
2.2 Simpli ed Viscoelastic Analysis . . . . . . . . . . . . . . . . . . . . . 48
4
2.3 Simple Thermal Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . 49
2.4 Temperature Dependency of Viscoelastic Properties . . . . . . . . . . 51
2.5 Materials with Pronounced Softening Temperature . . . . . . . . . . 53
2.6 Solution for Series of Simple Thermal Cycles . . . . . . . . . . . . . . 56
2.7 Analogy Between Thermal and Residual Stresses . . . . . . . . . . . . 58
3 Temperature Solutions for a Convectively Cooled Plate 60
3.1 Determination of Temperature Fields . . . . . . . . . . . . . . . . . . 60
3.1.1 Equation of the Heat Conduction . . . . . . . . . . . . . . . . 61
3.1.2 Thermal Boundary Conditions . . . . . . . . . . . . . . . . . . 63
3.2 One-Dimensional Analytical Temperature Solution . . . . . . . . . . . 63
3.3 Illustration of the Analytical Solution . . . . . . . . . . . . . . . . . . 69
3.3.1 Symmetrical Temperature Solution . . . . . . . . . . . . . . . 69
3.3.2 Asymmetrical Temperature Solution . . . . . . . . . . . . . . 72
3.4 \ln-cos" Residual Temperature Solution . . . . . . . . . . . . . . . . . 73
3.4.1 Illustration of the \ln-cos" Solution . . . . . . . . . . . . . . . 76
3.4.2 Application of the \ln-cos" Solution . . . . . . . . . . . . . . . 79
4 The Residual Temperature Concept 82
4.1 Frozen-in Strains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
4.2 Frozen-in Temperature Gradients { Non-Conservative Vector Field . . 85
4.3 De nition of the Integration Path . . . . . . . . . . . . . . . . . . . . 87
4.4 Features of the Orthogonal Lines . . . . . . . . . . . . . . . . . . . . 89
4.4.1 Orthogonal Lines at Symmetry Planes . . . . . . . . . . . . . 89
4.4.2 Radially Distributed Orthogonal Lines . . . . . . . . . . . . . 90
4.4.3 Singular Points in the \Hot" Region . . . . . . . . . . . . . . 91
4.4.4 The \Weld" Line . . . . . . . . . . . . . . . . . . . . . . . . . 92
5 Finite Volume Residual Stress Analysis 95

5
5.1 Governing and Constitutive Equations . . . . . . . . . . . . . . . . . 95
5.2 Fundamentals of the Finite Volume Method . . . . . . . . . . . . . . 100
5.3 Discretised Forms of the Governing Equations . . . . . . . . . . . . . 103
5.4 The Iterative Segregated Solution Algorithm . . . . . . . . . . . . . . 108
5.5 The Orthogonal Line Search Procedure . . . . . . . . . . . . . . . . . 113
5.5.1 Cell Face Intersection Points . . . . . . . . . . . . . . . . . . . 113
5.5.2 The Direction Vector b for the Cell . . . . . . . . . . . . . . . 116
5.5.3 Numerical Determination of the Weld Line . . . . . . . . . . . 117
5.5.4 Numerical Integration along the Orthogonal Lines . . . . . . . 121
5.5.5 Optimisation of the Numerical Procedure . . . . . . . . . . . . 122
6 Validation of the Numerical Algorithm 125
6.1 Validation of the Orthogonal Lines Procedure . . . . . . . . . . . . . 125
6.1.1 Orthogonal Lines for an Assumed Solidi cation
Time Function . . . . . . . . . . . . . . . . . . . . . . . . . . 126
6.1.2 Elliptic Orthogonal Lines . . . . . . . . . . . . . . . . . . . . . 129
6.2 Validation of the \ln-cos" Residual Temperature/Stress Solution . . . 130
6.3 Numerical Studies of Residual Stresses in a Polymeric Plate . . . . . 134
6.3.1 Symmetrical Cooling Examples . . . . . . . . . . . . . . . . . 136
6.3.2 Asymmetrical Cooling Examples . . . . . . . . . . . . . . . . 142
6.3.3 Connection between Symmetrical and Asymmetrical Residual
Stress Pro les . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
6.3.4 Curved Plate Subjected to External Bending . . . . . . . . . . 151
7 Experimental Validation of the Model 154
7.1 Description of the Industrial Casting Process . . . . . . . . . . . . . . 154
7.1.1 Polymerisation Process . . . . . . . . . . . . . . . . . . . . . . 156
7.1.2 The Material at Elevated Temperatures . . . . . . . . . . . . . 157

6
7.1.3 The Material in the Cooling Zone . . . . . . . . . . . . . . . . 158
7.1.4 Properties of the Cast Sheet . . . . . . . . . . . . . . . . . . . 159
7.2 Description of the Cooling Experiment . . . . . . . . . . . . . . . . . 160
7.2.1 Cooling Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
7.2.2 Temperature Measurements . . . . . . . . . . . . . . . . . . . 161
7.2.3 Heat Transfer Conditions . . . . . . . . . . . . . . . . . . . . . 163
7.2.4 Distortions of the Plate . . . . . . . . . . . . . . . . . . . . . . 164
7.2.5 Annealing Pro les . . . . . . . . . . . . . . . . . . . . . . . . 165
7.3 Cooling Experiments with Filled PMMA Plates . . . . . . . . . . . . 166
7.3.1 Thermocouple Measurement Results . . . . . . . . . . . . . . 168
7.3.2 Finite Volume Heat Transfer Analysis . . . . . . . . . . . . . . 168
7.3.3 Finite Volume Residual Stress Analysis . . . . . . . . . . . . . 171
7.3.4 Numerical Predictions vs. Experiment . . . . . . . . . . . . . 173
8 Applications of the Residual Stress Model 179
8.1 Layer Removal Method . . . . . . . . . . . . . . . . . . . . . . . . . . 179
8.1.1 Theoretical Description of the Method . . . . . . . . . . . . . 180
8.1.2 Curvature Functions . . . . . . . . . . . . . . . . . . . . . . . 184
8.1.3 The Reverse Problem . . . . . . . . . . . . . . . . . . . . . . . 185
8.2 Cooling a Cast Plate in Multiple Cooling Zones . . . . . . . . . . . . 186
8.2.1 Finite Volume Cooling Simulation . . . . . . . . . . . . . . . . 187
8.3 Finite Volume Layer Removal Simulations . . . . . . . . . . . . . . . 190
8.3.1 Original Residual Stress Pro les . . . . . . . . . . . . . . . . . 190
8.3.2 Residual Stresses During Layer Removals . . . . . . . . . . . . 192
8.3.3 Validation of the Finite Volume Layer Removal Model . . . . 197
9 Summary and Conclusions 200

Appendix A Measurement of Mechanical Properties 206

7
Appendix B Measuring the Speci c Heat Capacity 212

Bibliography 216

8
List of Figures
1.1 Residual stress principle { asymmetrical frozen-in strains. . . . . . . . 23
1.2 Residual stress principle { symmetrical frozen-in strains. . . . . . . . 25
2.1 Orientation of axes in an in nite plate. . . . . . . . . . . . . . . . . . 43
2.2 Radius of curvature of the plate under thermal load T (z). . . . . . . . 45
2.3 Decomposition of the temperature pro le T (z) into its constituent
components. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
2.4 Relaxation modulus curves. . . . . . . . . . . . . . . . . . . . . . . . 48
2.5 Simple thermal cycle. . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
2.6 Residual stresses in the plate exposed to a long lasting simple cycle. . 51
2.7 Family of stress relaxation curves for a thermorheologically simple
material. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
2.8 Solidi cation front in a plate with the temperature pro le T (z; t). . . 54
2.9 Simulation of the continuous cooling process as a series of simple cycles. 57
3.1 The distribution of the roots n of the eigen-equation. . . . . . . . . 65
3.2 Mathcad implementation procedure for determining roots n of the
eigen-equation for the convective cooling temperature solution. . . . . 70
3.3 Temperature histories in the symmetrically cooled plate from the ini-
tial uniform temperature. . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.4 Temperature histories in the asymmetrically cooled plate. . . . . . . . 73

9
3.5 Temperature histories and the residual temperature distribution for
symmetrically cooled plate. . . . . . . . . . . . . . . . . . . . . . . . 77
3.6 Temperature histories and the residual temperature distribution for
asymmetrically cooled plate. . . . . . . . . . . . . . . . . . . . . . . . 79
3.7 Comparison of the full and simpli ed temperature solution for an
arbitrary symmetrical convective cooling case. . . . . . . . . . . . . . 80
4.1 Formation of the frozen-in temperature gradient eld. . . . . . . . . . 86
4.2 Integration lines orthogonal to solidi cation fronts. . . . . . . . . . . 88
4.3 Orthogonal lines at symmetry planes. . . . . . . . . . . . . . . . . . . 90
4.4 Cooling example { symmetrical cooling from all sides. . . . . . . . . . 91
4.5 1D Cooling example { intensive cooling from the top surface. . . . . . 92
4.6 Arbitrary cooling example. . . . . . . . . . . . . . . . . . . . . . . . . 93
4.7 Weld line in the cooling domain. . . . . . . . . . . . . . . . . . . . . . 94
5.1 A control volume of an arbitrary shape. . . . . . . . . . . . . . . . . . 101
5.2 Orthogonal and non-orthogonal cell face topologies. . . . . . . . . . . 105
5.3 Flow diagram for solving non-linear energy equation system. . . . . . 109
5.4 Flow diagram for solving decoupled momentum equation system. . . . 110
5.5 Cantilever beam in bending { problem with the convergence. . . . . . 112
5.6 The straight line in 3-D intersecting with the face plane. . . . . . . . 114
5.7 Orthogonal line tracking procedure. . . . . . . . . . . . . . . . . . . . 115
5.8 Second-order accurate line search procedure. . . . . . . . . . . . . . . 117
5.9 Orthogonal lines determined for various direction elds. . . . . . . . . 118
5.10 Determining the weld line. . . . . . . . . . . . . . . . . . . . . . . . . 119
5.11 Orthogonal line terminating on the weld line. . . . . . . . . . . . . . 120
5.12 Numerical integration along the orthogonal line. . . . . . . . . . . . . 122
5.13 Optimised orthogonal line search procedure. . . . . . . . . . . . . . . 123

10
6.1 Solidi cation fronts and orthogonal lines for the assumed solidi cation
time function. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
6.2 Elliptic orthogonal line. . . . . . . . . . . . . . . . . . . . . . . . . . . 129
6.3 Temperature histories for the \ln-cos" validation case. . . . . . . . . . 132
6.4 Residual temperature and residual stresses for the \ln-cos" validation
case. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
6.5 Geometry of the at plate. . . . . . . . . . . . . . . . . . . . . . . . . 134
6.6 Classi cation of the cooling cases. . . . . . . . . . . . . . . . . . . . . 135
6.7 Results of the FV calculation for the 1-D symmetrical cooling case. . 138
6.8 Results of the FV calculation for the 2-D symmetrical cooling case. . 139
6.9 Temperature histories at the vertical symmetry A-A for symmetrical
cooling cases. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
6.10 Residual temperature and normal residual stress component xx at
the vertical symmetry A-A for symmetrical cooling cases. . . . . . . . 141
6.11 Results of the FV calculation for the 1-D asymmetrical cooling case. . 143
6.12 Results of the FV calculation for the 2-D asymmetrical cooling case. . 144
6.13 Temperature histories at the vertical symmetry plane A-A for asym-
metrical cooling cases. . . . . . . . . . . . . . . . . . . . . . . . . . . 146
6.14 Residual temperature and normal residual stress component xx at
the vertical symmetry plane A-A for the asymmetrical cooling cases. . 146
6.15 Decomposition of the symmetrical residual stresses into the bending
and asymmetrical pro les. . . . . . . . . . . . . . . . . . . . . . . . . 149
6.16 Curved plate subjected to external bending moments. . . . . . . . . . 152
7.1 Schematic diagram of a typical continuous casting production line. . . 155
7.2 Schematics of the cooling rig used for experiments. . . . . . . . . . . 161
7.3 Polymeric plate subjected to cooling from ve cooling nozzles. . . . . 162
7.4 Variation of local heat transfer coeÆcients for arrays of slot nozzles. . 164
11
7.5 Comparison of temperature histories at characteristic locations on the
plate. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
7.6 Heat transfer coeÆcient distribution on plate cooling surface. . . . . . 169
7.7 Results of the FV frozen-in strain calculation for 5 nozzle cooling
experiment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
7.8 Results of the 3-D FV residual stress calculation for 5 nozzle cooling
experiment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
7.9 Deformed pro les determined experimentally for 5 nozzle cooling ex-
periment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
7.10 Experimental and numerically predicted deformed pro les at both
central cross-sections for 5 nozzle cooling experiment. . . . . . . . . . 175
7.11 Deformed pro les determined experimentally for 1 nozzle cooling case.177
7.12 Experimental and numerically predicted deformed pro les at both
central cross-sections for 1 nozzle cooling experiment. . . . . . . . . . 178
8.1 Geometry of the plate element with the layers to be removed. . . . . 181
8.2 Calculated temperature histories for a single and multiple cooling
zones test simulations. . . . . . . . . . . . . . . . . . . . . . . . . . . 188
8.3 Residual temperature elds for single and multiple cooling zones cases.189
8.4 Residual stresses for single and multiple cooling zones cases. . . . . . 191
8.5 Net-forces and net-moments contained within the plate as a conse-
quence of residual stresses. . . . . . . . . . . . . . . . . . . . . . . . . 192
8.6 Redistribution of residual stresses xx in the fast cooled plate sub-
jected to 8 successive layer removals. . . . . . . . . . . . . . . . . . . 194
8.7 Redistribution of residual stresses xx in the slowly cooled plate sub-
jected to 8 successive layer removals. . . . . . . . . . . . . . . . . . . 194
8.8 Results of the FV layer removal calculations for the fast and the slow
cooling cases. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
12
8.9 The layer removal curvature pro les and corresponding residual stresses
for the fast and slowly cooled plates. . . . . . . . . . . . . . . . . . . 197
A.1 Schematic representation of the experimental setup used for the uni-
axial tensile tests. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
A.2 Temperature dependence of Young's modulus for lled PMMA. . . . 208
A.3 Visco-elastic ow of the material at the elevated temperature. . . . . 209
A.4 Isothermal non-dimensional relaxation curves. . . . . . . . . . . . . . 210
B.1 Typical DSC output signals and predetermined temperature variation.213
B.2 Measured heat ow signals using the DSC. . . . . . . . . . . . . . . . 214
B.3 Temperature dependent speci c heat capacity. . . . . . . . . . . . . . 215

13
List of Tables
3.1 Material properties and cooling parameters for the symmetrical cooling. 69
3.2 Auxiliary constants for the temperature solution equation. . . . . . . 70
6.1 Thermal, mechanical and cooling properties for \ln-cos" validation
cooling case. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
8.1 Thermal, mechanical and cooling properties for multiple cooling zone
nite volume simulations. . . . . . . . . . . . . . . . . . . . . . . . . 187
A.1 Temperature-dependent coeÆcients for the non-dimensional relax-
ation modulus hyperbolic t. . . . . . . . . . . . . . . . . . . . . . . . 211

14
Abbreviations
CMM Coordinate Measuring Machine
CZ Cooling Zone
DSC Di erential Scanning Calorimetery
FE Finite Element
FV Finite Volume
LRM Layer Removal Method
LSQ Least Squares
PMMA Polymethyl methacrylate

15
Nomenclature

aK ; K Discretisation coeÆcients
aP ; P Central discretisation coeÆcients
b; Discretisation Source terms
c Speci c heat capacity
e(t) Strain history
h Heat transfer coeÆcient
hV Volumetric heat source
k Thermal conductivity
m Mass
m Time step counter
qb Boundary heat ux
t Time
tsol Solidi cation time
x; y; z Chartesian coordinates
x; y; z Solidi cation front coordinates
z0 Initial poisiton of the top surface
z1 Current poisiton of the top surface
A; C; D Arbitrary constants for the slab temperature solution
An ; Cn ; Dn Constants in the in nite series solution
C Integration path
Cp Speci c heat capacity
E Young's modulus of elasticity
E (t) Relaxation modulus function
E1 Long term modulus
FP Resultant force in centroid point
Fx Net-force in the x-direction
F1 ; F2 Substitute functions
G(t) Relaxation modulus in shear
H Normalised heat transfer coeÆcient

16
I Moment of inertia
J (t) Creep compliance function
K Bulk modulus
L Half-thickness of the plate
MP Resultant moment in the centroid point
My Net-moment about the y-axis
S Surface bounding volume V
T Temperature
T Mean temperature
Tres Residual temperature
T res Mean residual temperature
Tb Temperature of the boundary surface
Tc Cooling temperature
Tg Softening temperature
T1 Temperature of the surroundings
T0 Reference temperature
T Temperature di erence
V Volume
X0 ; Y0 Starting point on orthogonal line
Zn Substitute function
b Direction vector for orthogonal lines
bV Direction vector in vertex point
fb Body force
I Identity tensor
k Direction vector for cross gradient
n Unit normal vector
nf Unit face normal vector
o Direction vector for orthogonal gradient
q Heat ux vector
r Position vector
rf Cell face position vector
rP Cell centre position vector
u Displacement vector
uB Boundary displacement vector
v Velocity vector
v Vector function
[a]; [ ] CoeÆcient matrices
[b]; [ ] Source vectors

17
CoeÆcient of thermal expansion
Nondimensional position of the point on the weld line
; n Roots of the eigen-equation
1 First root of the eigen-equation
Æmax Maximum warp
" Normal strain
"0 Mean strain
"xx; "yy ; "zz Normal strains
"xy ; "yz ; "xz Shear strains
" Strain tensor
"fr Frozen-in strain increment
 Thermal di usivity
 Lame elastic constant
 Nondimensional intersection point parameter
 Lame constant (shear modulus)
 Poisson's ratio
 Reduced time
 Density
 Radius of curvature
 Normal stress
 ( t) Stress history
res Residual stress
xx ; yy ; zz Normal stresses
B Bending stress
 Stress tensor
xy ; yz ; xz Shear stresses
x; y Curvatures
(T ) Time{temperature transposing function
 Arbitrary physical quantity
Potential function

18
Chapter 1
Introduction
The unique properties of polymeric materials have led to their widespread use in
many engineering applications and also in various domestic and public service ap-
pliances. Performance of the nal polymeric product is of the highest importance to
the user in the process of selecting the right product. Dimensional stability, dura-
bility, deformation behaviour and resistance to fracture are usually at the top of any
list of desirable properties for a new product. For polymer manufacturers it would
be highly advantageous to know these properties even before the product has been
actually produced. Unfortunately, a good design does not necessarily lead to a good
product since the performance of the nal polymeric product can be greatly a ected
by the manufacturing process. The work presented in this Thesis aims to describe
main aspects of the polymeric plate manufacture and the in uence these may have
on the behaviour on the nal plate product.

1.1 General Considerations


Ideally, material properties should be dominantly dependent on the chemical com-
position of the material, as this would allow a very simple classi cations of various
materials. However, this simple rule does not apply to most polymeric materials.
Namely, during manufacturing, polymers are subjected to several complex physical
19
Chapter 1 1.1. General Considerations
processes within a relatively short period of time. These physical processes have
a great impact on the material properties after manufacture. The most important
processes of this kind are polymerisation and solidi cation, processes of material
structuring and shaping of the product. Each physical process is described by a
number of processing parameters, e.g. time, temperature, pressure, velocity, etc.
The changes in any processing parameters will in uence, improve or worsen the nal
material behaviour. A product designer needs to relate the in uence of processing
parameters to the overall behaviour of the product.
Most polymeric products are manufactured at elevated temperatures by extrusion,
injection or blow moulding, or casting, and in the nal stage of processing the
material is cooled down to room temperature. At high temperatures the material
is in a softened state. This softened state is continuously transformed into a solid
state as the temperature in the system decreases. Solidi cation is the most dominant
physical process which sets many of the properties of the nal product. Solidi cation
is closely related to the thermal history of the system, which can be controlled by the
processing parameters, such as initial temperatures of the material, cooling rates,
temperature of a cooling medium, etc.
Knappe [1] identi ed the main material properties in uenced by the solidi cation
parameters as inhomogeneity, anisotropy and residual stresses. A material is inhomo-
geneous if its properties are non-uniformly distributed. Inhomogeneity is regularly
noticed after the solidi cation of semi-crystalline polymers. The size of crystals and
the volume fraction of the crystalline phase developed in such materials are highly
dependent on the cooling rates. The inner regions of the product, which experience
slower cooling regimes, will develop higher crystallinity, thus creating sti er zones.
Therefore, after the solidi cation is complete, the sti ness will vary through the
product.

20
Chapter 1 1.1. General Considerations
A material is anisotropic if its properties are direction dependent. In polymeric
materials the anisotropy arises from molecular orientation. During processing, the
polymer molecules can become orientated. Forces applied on the softened material
during the forming cause polymer molecules to partially align. If this is followed by
cooling, the cold solid polymer is left with a frozen-in molecular orientation. Usually
molecule orientation follows the ow patterns or the direction of any external force,
e.g. stretch force.

Residual stresses are induced into the material during non-uniform solidi cation.
Throughout the processing, the material is subjected to a non-uniform temperature
distributions and thermal expansions. Thermal expansions result in stresses in the
material via the modulus of the material. Due to the low modulus in the softened
state, the material can sustain hardly any deviatoric stresses arising from thermal
expansions. The majority of deviatoric stresses vanishes from the system, although
temperature distributions are still present. These temperature distributions are
eventually removed from the system, but only after the solidi cation is complete.
While in the solid state, the material can resist any removal of the \remaining"
temperature gradients, thus inducing residual stresses in the system.
All three e ects usually accompany polymer solidi cation and can cause problems
at di erent stages in the exploitation of the product. The inhomogeneity and
anisotropy cause design diÆculties of the nal product. The description of the
product's deformation behaviour is now considerably more complex, since material
properties are varying from point to point. Most standard stress analysis expres-
sions are usually derived using constant material properties. Standard compliance
functions, relating external loads and deformations in the material are no longer ad-
equate and will have limited applicability for design. The e ects of residual stresses
is also very important. Residual stresses a ect not only the response of the product

21
Chapter 1 1.1. General Considerations
to external loads, but they in uence the dimensions of the product itself. With
residual stresses accumulated in the material, the desired shape and dimensions of
the nal product are often not achieved. A polymeric product is usually tted into
a larger structure; if its build-in dimensions fall outside the prescribed tolerances,
additional machining is required, an expensive and time-consuming operation which
should be avoided if at all possible. In such cases, the new processing parameters
should be sought. The a-priori determination of all processing induced e ects would
be highly bene cial to designers. Of all processing-induced e ects, residual stresses
frequently cause the biggest concern in the polymeric plate manufacturing.
Residual stresses are mostly associated with composite materials. A composite
material combines two or more dissimilar materials together. Almost any composite
material, produced at elevated temperatures, and used at ambient temperature,
contains residual stresses. Residual stresses in such materials are easily attributed
to the di erences in coeÆcients of thermal expansion of the constituent materials.
During thermal processing, full thermal contractions of constituent materials are
prevented, thus generating permanent residual stresses. In case of bre-reinforced
plastics, the polymeric matrix tends to contract more then the embedded bres,
leaving the polymer matrix with tensile stresses and bres in the state of permanent
compression [2]. These strains can be considered as \frozen-in" the material.
A homogeneous polymeric material also contains residual stresses after the thermal
processing. Here, the generation mechanism of residual stresses is less obvious due to
a single coeÆcient of thermal expansion. The simplistic process of freezing-in strains
into a homogeneous material is illustrated in Figure 1.1. Consider two thin layers of
the same material heated up from room temperature T0 to a di erent, but uniform
temperatures T1 and T2 respectively (Figures 1.1a, b). Both layers are volumetrically
deformed, with strains proportional to applied uniform temperatures. The layers

22
Chapter 1 1.1. General Considerations

Figure 1.1: Residual stress principle { asymmetrical frozen-in strains. a) layers at


room temperature, b) layers at elevated temperatures, c) layers stacked together, d)
external moment applied to prevent visible distortion, = tension, = compression,
e) deformation due to residual stresses.
are now joined together (Figure 1.1c). After cooling to the same temperature T0, the
layer which has been at the higher temperature (T2) tends to contract more, but its
deformation is prevented by the other layer. This will produce tensile stresses in it.
The layer which was initially at the lower temperature (T1) will be in compression.
The internal equilibrium in material layers is disturbed and an external moment
M must be applied to keep the new geometry at (Figure 1.1d). If the external
moment M is removed, the internal stresses will redistribute, and the new geometry
curves (Figure 1.1e). The permanent residual stresses are induced in the material
23
Chapter 1 1.1. General Considerations
as a result of thermal e ects. The continuous process, analogue to the above, takes
place when polymeric material is subjected to a non-uniform solidi cation process.
The distortions due to residual stresses are obvious in the production of polymeric
plates: it is very diÆcult to achieve a at sheet after the solidi cation process.
Usually, the plate is curved in some way after cooling. The more the solidi cation
process is symmetrical, the more at the sheet will be obtained after cooling, as
symmetrical solidi cation produces symmetrical frozen-in strain distributions. The
analogue simplistic process for the symmetrical solidi cation can be schematically
explained using Figure 1.2. The resulting symmetrical residual stresses (Figure
1.2d) are not producing a bending moment responsible for curving the geometry.
Although in such case there is no obvious macroscopic distortion, the geometry is
heavily internally strained. If in such case the internal equilibrium of the plate is dis-
turbed by sectioning the plate, the strain is no longer symmetrical and macroscopic
deformation occur. The deformation is proportional to the magnitude of stresses in
the removed section and can be controlled by the cooling rate.
Warpage and conditions which lead to warpage should be avoided when producing
polymeric plates. The two main guidelines for the polymeric plate manufacture are
therefore: symmetrical cooling and lower cooling rates. The latter condition is in
contradiction with the economic demands. A predictive model which relates the
manufacturing process parameters with the residual stresses and associated distor-
tions would be a most welcome tool for plate manufacturers, as such model would
help obtain the optimum balance between technical requirements and economic de-
mands.

24
Chapter 1 1.2. Computational Methods

Figure 1.2: Residual stress principle { symmetrical frozen-in strains. a) layers at


room temperature, b) layers at elevated temperatures, c) layers stacked together,
d) no visible distortion due to symmetrical residual stresses,  = tension, =
compression.
1.2 Computational Methods
The use of numerical methods and powerful computers has evolved in the past few
decades. Numerical methods are widely accepted in modern engineering practice [3].
Nowadays, a number of commercial numerical packages is available and capable of
solving various physical problems in uid mechanics, solid mechanics, heat transfer,
electromagnetism, etc. For example, ABAQUS [4] and ANSYS [5] are well suited
for solving problems of solid mechanics, FLUENT [6] and FOAM [7] solve uid ow
problems, C-MOLD [8] and PHYSICA [9] address problems of phase change, casting,
etc. In all these packages, the basic physical principles, governing various physical
properties, are simulated using the appropriate numerical methods. Governing prin-
25
Chapter 1 1.2. Computational Methods
ciples can be expressed with the systems of partial di erential equations, for which
the analytical solution rarely exist. Even the idealised one-dimensional problems are
often analytically not solvable. For a limited number of situations where analytical
solution does exist, the form of the solution is often mathematically very complex.
The numerical methods transform complex equations of the mathematical model
into systems of algebraic equations that can be solved on the computers. The
process starts by splitting the material domain into smaller elements to create a
computational mesh. Similarly, the time domain is split into a number of time steps.
Further, the governing equations are integrated over each element by employing
simple approximation functions for each unknown variable. This process is known
as discretisation, and is usually performed using the Finite Element (FE) or Finite
Volume (FV) discretisation techniques. While the FE method is more orientated
towards resolving problems in solid mechanics, the FV method is preferred method
in the computational uid dynamics community, mainly due to a high level of the
conservativeness that it possesses, and its iterative nature required for solving non-
linear problems [10]. Being conservative, the FV method is becoming increasingly
popular in the solid mechanics for solving dynamic and highly non-linear problems,
e.g. dynamic fracture mechanics [11], etc. Consequently, complex problems of a
multi-physics, with the strong interactions between solids and uids can be uni ed
into a single FV framework [12].
After the discretisation, the systems of algebraic equations are formed. These are
usually very large, containing thousands of algebraic equations with thousands of
unknown variables. Such large systems are solved using iterative equation solvers.
Development of all necessary tools used in the numerical methods arose as a sep-
arate multi-disciplinary eld of research. However, before attempting to solve any
problem, a full understanding of all aspects of the physical problem is required.

26
Chapter 1 1.2. Computational Methods
The commercial numerical packages cannot always incorporate all the speci c de-
tails of the problem of interest. The ability to incorporate user-de ned procedures
into the commercial packages would make them even more powerful to the users.
User-de ned procedures can customise, ne-tune and optimise the solution algo-
rithm for a very speci c physical problem, or for a speci c geometry. The few areas
where the user can enhance the overall eÆciency of the numerical procedure include
the speci c material behaviour, the new correlations between the dependent vari-
ables, possible and well justi ed simpli cations, the use of analytical solutions etc.
However, developing these procedures may be unpopular with the users, since it
often requires a comprehensive and detailed knowledge of the numerical package on
a much deeper level of implementation. Rarely does the user have access to these
parts of the program. But, if possible, the user-de ned procedures may be bene cial
to use.
A mathematical model for prediction of residual stresses in polymeric plates has
to incorporate the dominant physical phenomena which are taking place during
processing. At elevated temperatures, viscoelastic ow or rubber-like material be-
haviour are dominant. In any case, the material is in the softened state. This will
result in fast stress relaxation processes. It can be assumed that at high tempera-
ture the material is virtually stress-free due to these relaxation processes. At room
temperature, the polymeric material can be characterised as an elastic solid. Any
deformation is accompanied by stresses corresponding to the elastic stress-strain
curve, although time dependent stress relaxation is still present, but at a lower rate.
The frozen-in strains, present in the material after processing, are a consequence
of inelastic e ects which took place between the high and low temperature events,
i.e. during the non-uniform solidi cation. The viscoelastic material behaviour and
previous related work relevant to residual stress problem in polymeric plates will be
discussed below.
27
Chapter 1 1.3. Basic Principles of Linear Viscoelasticity
1.3 Basic Principles of Linear Viscoelasticity
The basic theory of linear viscoelasticity is systematically presented in [13]. Vis-
coelastic materials incorporate characteristics of both elastic (Hookean) solids and
viscous (Newtonian) uids. The main di erence in the deformation behaviour of
these two material groups is the ability to restore the energy supplied by external
loads during deformation. While elastic solids accumulate this energy as a strain
energy, which is fully recoverable upon removal of external loads, viscous uids dis-
sipate the supplied energy through irreversible ow of the material. Any viscoelastic
material can easily display both of these extremes of the deformation spectrum. At
short times the response is closer to elastic solids, since no time was allowed for the
viscoelastic ow, while at longer times, the energy becomes dissipated to a great
extent.
In linear viscoelasticity the Boltzmann's superposition principle [14] is employed to
determine the viscoelastic response to an arbitrary external load. Mathematically,
this leads to a constitutive relation for the material in the form of a convolution
integral. The convolution integral can be expressed in several di erent forms, the
most widely employed are formulations with the time dependent relaxation modulus
G(t) and creep compliance J (t) functions [15]:
Zt
de( )
Stress relaxation form:  (t) = G(t  )
d
d; (1.1)
0

Zt
d ( )
Creep compliance form: e(t) = J (t  )
d
d: (1.2)
0

Here,  and e are stress and strain respectively, and t and  are time variables.
The convolution integrals are often referred to as heredity integrals, since a stress
at any time is dependent upon the entire strain history e(t), and similarly for the
28
Chapter 1 1.3. Basic Principles of Linear Viscoelasticity
stress history (t). In any arbitrary loading case, both creep and stress relaxation
processes are taking place simultaneously.
Stress relaxation and creep are the basic phenomenological viscoelastic responses of
the material and both can easily be measured experimentally. In the stress relaxation
experiment the material is subjected to an \instantaneous" strain e0. The strain is
maintained constant throughout the experiment, while decaying stress is recorded
as a function of time (t). In the creep experiment the variables are reversed. The
stress 0 is imposed onto the specimen and is kept constant, while the strain is
measured as a function of time e(t). From this basic responses, the stress relaxation
and creep compliance functions, required for Equations (1.1) and (1.2), are obtained:
 (t) e(t)
G(t) = and J (t) = : (1.3)
e0 0
A unique relation between J (t) and G(t) exists; however they are not simply re-
ciprocal functions. The expressions which contain convolution integrals can be
conveniently manipulated mathematically only after transforming the constituent
functions using the Laplace operator [16]. By combining the Laplace transforms of
expressions (1.1) and (1.2), the following relationship between stress relaxation and
creep compliance functions can be obtained [17]:
Zt Zt
dJ ( ) dG( )
G(t  ) d = J (t  ) d = 1: (1.4)
d d
0 0

Therefore, for any linear viscoelastic material, the creep compliance J (t) can be
calculated from the stress relaxation response G(t), and is uniquely determined.
The convolution integrals are mathematically complex to solve for arbitrary loading
cases. But simpli ed mechanical models can be used as an approximation for a
viscoelastic material. The simpli ed models consist of a number of linear spring and

29
Chapter 1 1.3. Basic Principles of Linear Viscoelasticity
viscous dashpot elements. A spring element of constant sti ness G portrays the be-
haviour of a Hookean material, while a dashpot with a constant viscosity  describes
Newtonian uids. Di erent numbers and combinations of these one-dimensional
mechanical elements lead to di erential equations which describe the deformation
behaviour of a polymeric material. The simplest to use are two-component models,
consisting of one spring and one dashpot element. A combination of a spring and
dashpot in series is known as the Maxwell model, whereas the parallel combination
is the Kelvin or Voigt model [15]. Unfortunately, both this simple models fail to
describe one of two basic viscoelastic responses, either stress relaxation or creep.
Additional elements must be included to produce the response closer to a real poly-
meric material. If a second spring is added in parallel with the Maxwell element,
the more realistic response is obtained. This model is known as a standard linear
solid [15]. However, no combination of springs and dashpots will satisfy the Equa-
tion (1.4), and this automatically induces errors into the mathematical model [17].
Therefore, the convolution integral is the preferred form for describing the behaviour
of linear viscoelastic materials.
Up to this point, the temperature did not feature at all in the description of vis-
coelastic material behaviour. Intuitively, we expect the stress relaxation and creep
processes to be accelerated at higher temperatures. Relaxation and creep curves
can be measured experimentally by maintaining isothermal conditions. After a se-
ries of stress relaxation tests, a number of stress relaxation curves are obtained, each
valid only at a constant temperature. If the temperature during the experiment is
not constant, the material will still be relaxing. The aim here is to predict this
response from the measured isothermal relaxation curves. A simple switch from one
relaxation curve to another, at a speci ed time, is physically not plausible.
Consider two stress relaxation tests in which two identical specimens are subjected

30
Chapter 1 1.3. Basic Principles of Linear Viscoelasticity
to the same strain level e0. In the rst test, the material is relaxing at a low
temperature T1 and stresses are dropping slowly according to the low temperature
relaxation curve. After a long time, the stress is still present in the specimen.
In the other test, the material is initially maintained at a high temperature T2 ,
and afterwards the temperature is lowered to T1. While at high temperature, the
stress has relaxed to a great extent, it has reached a much lower level then in the
low temperature relaxation test. By lowering the temperature to T1 , the stress in
the second test is not expected to rise back to the level determined by the low
temperature relaxation curve. It is expected that the stress continues relaxing,
but at a lower rate. This ever-decreasing stress during the relaxation test is the
basis for the time{temperature superposition principle [15, 18], a principle used for
determining viscoelastic response of materials subjected to temperature changes.
In the time{temperature superposition principle, the relaxation process at a high
temperature T2 is equivalent to the relaxation process at a low temperature T1 ,
but at much longer relaxation times. For \thermo-rheologically simple" materials
[19] the relaxation process at high temperature is accelerated by a factor which
is constant in a wide range of relaxation times. This constant factor is evident if
the relaxation curves are plotted in the diagram with the log (time) abscissa. The
relaxation curves for di erent temperatures have the same shape in the diagram,
they are just translated along the log (time) axis. This translation factor is known
as a \shift factor" and can be determined from experimental traces, one factor for
each temperature.
The previous example illustrates that the stress during the relaxation test is depen-
dent on the complete temperature history to which the material was subjected. The
role of temperature here is twofold, it in uences the viscoelastic material properties,
and also causes thermal stresses via the coeÆcient of thermal expansion. This de-

31
Chapter 1 1.4. Previous and Related Studies
pendency must be incorporated into the convolution integral constitutive law, which
would allow the determination of material response to any arbitrary mechanical and
thermal conditions. The detailed mathematical description of the time{temperature
superposition principle will be presented in the Chapter 2.

1.4 Previous and Related Studies


The problem of residual stresses in viscoelastic materials has received considerable
attention in the past. In the early years of the last century, Adams and Williamson
[20] set the fundamentals in the area of annealing of inorganic glasses. Substantial
economic losses were incurred by breakage of glass attributed to residual stresses.
The sensitivity to residual stresses was even more pronounced with optical glass,
where geometrical distortions due to residual stresses spoiled the optical charac-
teristics of the glass. Although viscoelastic theory was not known at the time,
they recognised that stress relaxation at elevated temperatures was the main reason
for generation of residual stresses. Adams and Willaimson [20] claimed that if a
stress-free glass plate that contains temperature gradients, is brought to a uniform
temperature, the residual stresses in it are opposite to thermal stresses which would
be produced by the same temperature gradient. The analogy between thermal and
residual stresses will prove a useful concept in residual stress works that followed.
Full mathematical description of viscoelastic processes were published in 1960s,
again inspired by the glass industry. The mathematical models were very com-
plex and rarely solvable even for the simplest one-dimensional situations. Lee et
al. [21] used the convolution integral to relate the deviatoric stresses and strains in
order to calculate the residual stresses in a glass plate. The dilatational response
was assumed to be fully elastic, with the bulk viscosity approaching in nity (the
bulk modulus does not decay in time). Temperature changes were accounted for
32
Chapter 1 1.4. Previous and Related Studies
via the time{temperature superposition principle. Very cumbersome mathematical
methods were employed to resolve the nal integral equations for residual stresses,
often involving Laplace transformations and their inverse. However, the analysis
was restricted to symmetrical one-dimensional solidi cation. In a later publication,
Lee and Rogers [22] used the measured relaxation functions to solve the transient
stresses for PMMA. Mathematical diÆculties arose in obtaining the inverse Laplace
transforms for some auxiliary functions used in the algorithm, which are based on ex-
perimentally measured relaxation curves. The simpli ed versions of these, based on
the Maxwell model, were resolved analytically. A very general and mathematically
well posed transient thermal stress solution was presented by Muki and Sternberg
[23]. In contrast to the analysis from Lee et al. [21, 22], they considered the general
viscoelastic responses in both shear and dilatation. The exact solution was again
mathematically unobtainable for measured material properties, these were obtained
only for simpli ed Maxwell and Kelvin type viscoelastic solids.
Mathematical complexity of the stress solutions also arises from transient thermal
solutions, which determine temperature distributions in the material. This solution
can be expressed analytically for a limited number of realistic situations and for
idealised thermal material properties. Carslaw and Jaeger [24] derived a number
of analytical solutions for slab geometries. They covered a wide range of thermal
conditions, slabs with uniform and non-uniform initial temperatures, surfaces sub-
jected to convective cooling, insulated surfaces, etc. In glass manufacture, with
temperatures typically higher then 500ÆC, substantial heat losses occur through ra-
diation e ects [25]. Such problems require numerical solutions for temperature, and
consequently numerical solutions for stresses. The solidi cation process for some
polymers additionally include the phase change processes, which do not happen in
glass manufacture. The temperature solution for such cases needs to account for
latent heat of the phase change. Numerical methods have been developed to resolve
33
Chapter 1 1.4. Previous and Related Studies
such thermal systems [26, 27], while analytical solutions for these are very rare and
complex.
The complexity of the exact solutions for tempering stress problems in glass resulted
in simpli ed models in subsequent publications. Although these models oversimpli-
ed the real material behaviour, they calculated remarkably accurate solutions for
practical purposes. In their formulation, Aggarwala and Saibel [28] used the Maxwell
model to describe glass, a model that was traditionally accepted to represent glass
behaviour quite well. The viscosity in their model was a discontinuous function of
temperature. The coeÆcient of viscosity was set to zero at temperatures above the
critical temperatures, and in nity below it. Their model was similar to the "instant-
freezing" model, rst introduced by Bartenev [29]. By de ning the glass forming
temperature Tg at which the viscoelastic properties change substantially, Bartenev
introduced a ctitious temperature function which correlates the residual stresses to
the thermal history in the glass plate. The parabolic residual stress pro les obtained,
with the compressive stresses on the surface and the inner tensile region, correlated
well with the experiment. The instant freezing model cannot predict the transient
thermal stresses but only the residual stresses at the end of cooling. The distribution
of thermal stresses during the glass cooling of a glass are somewhat di erent from
the parabolic residual stress pro le. The full evolution of thermal stresses for several
symmetrical glass cooling cases were reported by Narayanaswamy and Gardon [30].
The ndings from the analysis of inorganic glasses were later transposed to polymeric
materials, since measured residual stress pro les in polymers were also parabolic.
Williams [31] proposed the residual stress model for polymers, where viscoelastic
analysis includes the time{ temperature superposition principle, in which the shift
factor was discontinuous at the softening temperature Tg . Below the softening tem-
perature, the shift factor was unity and above it was in nity. This is another way

34
Chapter 1 1.4. Previous and Related Studies
of implementing the instant-freezing idea. In [31], Williams evaluated the residual
stresses for one-dimensional slab and pipe geometries. In the continuation of this
work, the algorithm was recently extended beyond the one-dimensional solidi cation
cases in [32, 33]. Other polymer researchers used similar ideas in their residual stress
algorithms. Miyano et al. [34] analysed residual stresses in quenched thermosetting
beams. They qualitatively described the growth mechanism of residual stresses us-
ing the Maxwell viscoelastic mechanical model, in which dashpot was free to move
at temperatures above the glass transition temperature Tg , and was xed below Tg .
In the quantitative viscoelastic analysis they used the time{temperature superpo-
sition principle, where the master relaxation curve was represented mathematically
by the Prony series, and the shift factor function was obtained from the theoretical
Arrhenius activation energy equation, as can be found in [15]. From experimental
and numerical analysis, Miyano et al. obtained parabolic residual stress pro les
for cases when quenching temperature was above Tg , and almost no stresses when
the quenching was performed from temperatures below the Tg . Similar observations
were reported for quenched epoxy plates by Maneschy et al. [35].
The Prony series representation became a popular method of describing mathemati-
cally the master relaxation curves, for both shear and bulk modulus [36]. The master
relaxation curve is a relaxation curve, valid for a single reference temperature over
a very long period of time, usually more then ten time decades in the logarithmic
time diagram. It is a very convenient representation for conventional materials since
numerous Prony series coeÆcients can be found in the literature. However, deter-
mination of these coeÆcients for a speci c material, and many variants of the same
material, is complex and time-consuming.
In the production of many polymeric products, the material is either polymerised
or cured just before the solidi cation process. Prior to polymerisation or curing,

35
Chapter 1 1.4. Previous and Related Studies
the material has virtually no mechanical sti ness and it is often in the form of
viscous liquid. As polymer chains grow, the material hardens, and stresses can
develop. The shrinkage which accompanies the hardening process can result in
stresses in the material, especially if the material is being moulded in closed moulds
and thus prevented from shrinking. For products that harden in an open geometry,
large hydrostatic stresses that can arise from the bulk modulus are avoided. A very
descriptive basic principle for hardening processes for epoxy resins are given by Skeist
[37]. The duration of the curing process determines the degree of material conversion,
which has a great impact on the viscoelastic properties of the material. Shimbo et
al. [38] investigated the relaxation behaviour of a thermosetting resin subjected to
di erent curing regimes. They found the reciprocation law of time{temperature{
degree of cure, which attempts to determine the relaxation properties of a resin
during curing. Similar to the time{temperature superposition principle, the master
relaxation curve can be shifted horizontally and vertically in the log (time) relaxation
diagram to account for curing e ects. The glass transition temperature Tg is also
a ected by the degree of cure. Martin and Adolf [39] also recognised the time-cure
superposition principle as a potential theoretical approach to describe viscoelastic
behaviour of polymers during cure. Residual stresses are also of concern in the
electronic industry where integrated circuits are often buried into epoxy resin blocks.
Semi-conductor components experience substantial compressive stresses during the
curing process due to shrinkage of the surrounding epoxy [40], but in the single
material systems these stresses are insigni cant.
Nowadays, with the ever developing computer technology, computing capabilities
have greatly risen. More sophisticated and larger problems in polymer engineer-
ing are solved numerically. Physical processes which can be simulated range from
the ow of molten polymer, solidi cation, morphology predictions, residual stresses
and distortions, etc. However, the fundamental physical principles did not evolve
36
Chapter 1 1.4. Previous and Related Studies
much from those described previously, they are just uni ed into a single numerical
framework which was not possible to solve in the past. Numerical simulation of
injection moulding process is receiving considerable interest today, as one of the
most important polymer processing operation. Chiang et al. [41, 42] presented a
uni ed modelling method for simulating lling and post lling stages of the injec-
tion moulding process. The cavity lling process is described by a general 3-D uid
ow equations, with 1-D and 2-D ow approximations employed where possible.
Thin parts are analysed as a planar 2-D geometries and thermally induced residual
stresses in there are calculated explicitly using the model from Muki and Sternberg
[23]. Although many predicted and experimentally measured results are in good
agreement [43], the numerical model has many empirical correlations, speci c for
the given mould geometry and the given polymer. Similar treatment for thermally
induced residual stresses in injection moulded thin shapes is reported by Chang et
al. [44], where thermal stresses are computed explicitly from the pressures and tem-
peratures obtained from the uid ow results. Another challenging process in the
plastic industry today is the numerical simulation of the plastic pipe manufacture.
Pittman and Farah [45] presented a comprehensive model for calculating residual
stresses and shrinkage of a semi-crystalline polymer pipe. Thermal stresses during
the cooling process are described by a generalised Maxwell model with tempera-
ture dependent parameters. Additionally, the development of crystallinity is linked
to the cooling process, which enables some predictions of morphology for the pipe
to be made. The predicted hoop residual stresses in the pipe wall are parabolic,
being compressive on the cooled boundaries. Compressive stresses are bene cial
on both pipe surfaces since they enhance the resistance to crack propagation from
defects that potentially exist on pipe surfaces. Broutman et al. [46] investigated
the in uence of surface compressive residual stresses, induced by quenching from a
temperature above the Tg , on the impact strength of polymers. Large compressive

37
Chapter 1 1.4. Previous and Related Studies
stresses on the surface suppress the craze formation and growth, which delay brittle
fracture in the material. This has a positive e ect in the exploitation of plastic pipes.
In the work by Akay and Ozden [47] on the injection moulded polycarbonate, the
presence of residual stresses signi cantly increased the toughness under static and
impact conditions, while other measured properties changed only slightly (elastic
modulus, yield strength, glass transition temperature).
Apart from theoretical models for prediction of residual stresses in polymers, con-
siderable e ort in the past has been directed towards developing methods for mea-
suring them. For sheet materials, the layer removal method (LRM) is employed
most widely. The pioneering work by Treuting and Read [48] explained the theo-
retical aspects of the layer removal method. Here, progressive uniform layers are
removed from the surface of the plate geometry, thus disturbing the residual stress
equilibrium. Consequential changes in the curvature can be measured and related to
stresses which existed in the material prior to layer removal. Although destructive
and very time-consuming, the layer removal method can produce reliable residual
stress pro les. Several layer removal measurement examples can be found in [49] and
[50]. In another variant of the layer removal method, the relieved strains during layer
removal can be measured by strain gauges [51], and related to the original residual
stresses. The layer removal method can also be employed for non-homogeneous ma-
terials, where material properties, like the elastic modulus and the Poisson's ratio,
vary through the thickness. Eijpe and Powell [52, 53] developed a modi ed layer
removal methods suited for composite materials. Composite plates may exhibit
very large deformations during layer removals and often twist due to residual shear
stresses. The main disadvantage of the layer removal method is that it is highly de-
structive. It also has limited applicability in cases of laterally non-uniform residual
stress elds, where the method produces an average residual stress eld in the whole
sample.
38
Chapter 1 1.4. Previous and Related Studies
Another popular and less destructive residual stress measurement method is a hole-
drilling method, which is also an ASTM standard method [54]. The hole drilling
method involves localised removal of the stressed material by drilling a small diam-
eter hole in the material. The strain relief in the surrounding material is measured
using a specialised three-element strain gauge rosette. Measured strains can be
correlated to the residual stresses in the material. The method can measure very
localised, through-thickness residual stresses. Unfortunately, it is very sensitive to
experimental errors, since strain relief on the material surface, at the location of the
strain gauge, is low. A systematic overview of many other residual stress measure-
ment methods can be found in [55].
In the concluding part of this review, the following summary can be made. Residual
stresses are always induced into visco-elastic materials during the thermal process-
ing. Most of the residual stresses are generated during the cooling from a tem-
perature above Tg , while material is going through non-uniform solidi cation. The
predictive residual stress models correlate thermal history to a mechanical response
of the material during processing. The e ect of residual stresses can sometimes
be bene cial in the exploitation of the product, but in many more cases residual
stresses should be avoided or minimised. Their distribution should be known in any
case. The sophisticated numerical programs can assist polymer manufacturers in
nding optimal processing parameters in order to manipulate the residual stresses
to desired e ect.

39
Chapter 1 1.5. Layout of the Thesis
1.5 Layout of the Thesis
The rest of the Thesis will be organised as follows:
In Chapter 2, the thermo-viscoelastic analysis is performed for the at plate geom-
etry. The analysis is focused on a one-dimensional symmetrical cooling case. As a
result, the frozen-in strain patterns are identi ed and the resulting residual stresses
are evaluated analytically. The frozen-in strains are a function of thermal history
during processing that are described with the residual temperature eld.
Chapter 3 describes the one-dimensional transient thermal solution for a slab sub-
jected to convective cooling. Detailed mathematical derivation of the problem is
given, and the solution is illustrated on two cooling examples: symmetrical and
asymmetrical cooling. The analytical solution for the residual temperature eld,
based on the approximate convective cooling thermal solution is derived and illus-
trated on the same examples.
The frozen-in strain principle from Chapter 2 is generalised in Chapter 4 to arbitrary
cooling/solidi cation cases. With the appropriate method for determining the resid-
ual temperature eld, the residual stress analysis is extended to multidimensional
analysis. The material responses in the same way to residual stresses, as to the
applied residual temperature eld. The theoretical aspects and important features
of the novel orthogonal line algorithm are discussed.
Chapter 5 gives an overview of the Finite Volume method, employed to solve the
equations governing energy and momentum during the numerical simulation of the
residual stress problem. Numerical aspects of the method, including discretisation
techniques and solution algorithms are outlined. Speci cally for the residual stress
problem, the solution procedure for determining orthogonal lines is adapted to the
nite volume mesh topology. Mathematical algorithms of all steps in the orthogonal
lines algorithm are given.
40
Chapter 1 1.5. Layout of the Thesis
In Chapter 6, the proposed Finite Volume solution procedure is applied to several
test cases. Initially, the numerical solution for orthogonal lines, the residual tem-
perature eld and residual stresses are veri ed against available analytical solutions.
The numerical validation is followed by simulations of four cooling examples relevant
to at plate manufacture: symmetrical and asymmetrical, one and two-dimensional
cooling cases. The results for the asymmetrical cooling cases are related to the
symmetrical cooling case results in a detailed theoretical and numerical discussion.
In Chapter 7 the frozen-in strain model is examined numerically and experimentally
for a speci c polymeric material. The cooling experiments are performed using the
lled PMMA plates, produced by the continuous casting process. The objective of
the experiment is to relate the cooling conditions in the plate to the deformations in
the plate after the experiment, which are attributed to residual stresses. The exper-
imentally measured thermal histories are later used to determine thermal conditions
for the 3-D numerical analysis. The numerically predicted pro les are compared
with the experimentally measured deformations.
Chapter 8 illustrates some practical situations which can be resolved using the pro-
posed numerical model. A single cooling zone from the industrial casting line can
be replaced by multiple cooling zones to reduce the residual stresses in the sheet.
Cooling times for each cooling zone and the resulting residual stresses are predicted
numerically. The layer removal technique is simulated numerically for sheets pro-
duced on a single and multiple cooling zone lines. The detailed theoretical and
numerical descriptions for the layer removal method are also presented.
Finally, Chapter 9 summarises the Thesis and o ers some conclusions and sugges-
tions for future research.

41
Chapter 2
Viscoelastic Stress Analysis in a
Flat Plate
The main reason for the presence of residual stresses in polymeric materials after
thermal processing are \frozen-in" strains. These strains give rise to stresses at
room temperature through the modulus of the solid material and remain \locked"
in the material with no ability to relax on their own. The objective of residual
stress analysis is to estimate the frozen-in strain distribution within the material
as a function of the thermal history the material went through during thermal
processing. In this Chapter, the equations for thermally induced residual stresses
will be developed.

2.1 Thermo-Elastic Analysis of a Flat Plate


The derivation starts with the equilibrium equations and the constitutive law for
a linear thermo-elastic solid. The compatibility relations are also employed as a
guarantee that no discontinuities are developed in the material during thermal pro-
cessing. This set of equations can be solved for stresses arising from mechanical
and thermal loads. When the material is free of external mechanical loads and
constraints, the stress distribution in the material is solely caused by the tempera-

42
Chapter 2 2.1. Thermo-Elastic Analysis of a Flat Plate
ture gradients in the material. Furthermore, at a uniform temperature the thermal
stress component vanishes, leaving the material in a stress-free state. Thermo-elastic
analysis as such cannot produce stresses in the material free of external loads and
constraints at a uniform temperature, i.e. residual stresses. Residual stresses in
the material are a consequence of anelastic e ects the material is going through
during thermal processing. A modi cation to the thermo-elastic analysis must be
introduced to accommodate these e ects.
The simplest case to analyse is a thin at plate, shown in Figure 2.1, in which the
Cartesian coordinates (x; y; z) will be used. The following analysis is based on the
previous work of Williams [31]. The origin of coordinate system may be taken at
any point in the central plane of the plate, with the z-direction perpendicular to the
central plane. The plate has the thickness of 2L, so that z varies from L to +L. The
plate is assumed to be free of surface tractions and body forces. Distribution of the
temperature in the plate is such that it varies only across its thickness: T = T (z),
leaving the plate free of temperature gradients in the other two directions (x; y)
(Figure 2.1).
z
z=+L

T=T(z)

x 2L z=0
z=-L

Figure 2.1: Orientation of axes in an in nite plate.

43
Chapter 2 2.1. Thermo-Elastic Analysis of a Flat Plate
From the stress-free conditions it follows zz = xz = yz = 0 on the top and
bottom surfaces. Under such loading, there exists a total rotational symmetry of all
dependent variables about the z-axis and therefore xx = yy = (z). These stress
components can also be considered as principal stresses, reducing the stress state
in any point in the plate to a point in the Mohr's diagram, which yields xy = 0.
The strain components can be treated in a similar manner, "xx = "yy = "(z) and
"xy = "yz = "xz = 0. Additionally, the compatibility equations give the following
di erential form of allowable variations in strains [56]:
@ 2 "xx @ 2 "zz
= @@x@z
"xz
2

@z 2
+ @x2 ;
@ 2 "yy @ 2 "zz
= @@y@z
"yz
2

@z 2
+ @y2 :

Both compatibility equations are reduced to:


@2"
@z 2
= 0: (2.1)
Hence, the normal strain " varies with the z only, the left side in Equation (2.1) is
the total di erential of ", and can be integrated twice giving linear variation of the
strain through thickness:
z
"(z ) = "0 + : (2.2)

Analogy with the plate bending theory identi es two integration constants "0 and
 as the average strain and the radius of curvature of the plate respectively (Figure
2.2). The general form of the Hooke's law for thermo-elastic solid:
" = [
1  ( +  )] + (T T )
xx xx yy zz 0
E
is simpli ed for this stress state and becomes:

"= (1  )+ (T T0 ): (2.3)
E
44
Chapter 2 2.1. Thermo-Elastic Analysis of a Flat Plate

z
T=T(z)

z=0
ρ

Figure 2.2: Radius of curvature of the plate under thermal load T (z).
The mechanical properties E and  in Equation (2.3) are the Young's modulus of
elasticity and the Poisson's ratio of the material. The last term in the Hooke's law
accounts for the thermal expansion of the material. The strain that corresponds
to zero stresses is no longer zero, but is increased to the value of (T T0 ) that
material would posses when allowed to expand freely. This new \referenced" state
depends on the coeÆcient of the thermal expansion and the di erence between
actual temperature T and a reference temperature T0. At T0 the material is in
undeformed state. This temperature is usually zero or the room temperature. Com-
bining Equation (2.2) with Equation (2.3), the stress distribution in the plate with
temperature distribution T (z) becomes:
 
E z
 (z ) = "0 + T (z ) : (2.4)
1  
It is desirable to express the remaining unknowns in Equation (2.4), "0 and , in
terms of known temperature eld T (z). The net-force and the net-moment calcu-
lated from the above stress pro le must be in the equilibrium with the forces and
moments applied to the plate boundary. Since there are no mechanical constraints

45
Chapter 2 2.1. Thermo-Elastic Analysis of a Flat Plate
on the plate, these quantities are zero, resulting in the following two equations:
Z+L Z+L

Net-force:  (z )dz = 0 ) "0 =
2L T (z )dz = T ; (2.5)
L L

Z+L +L

Net-moment:  (z )zdz = 0 ) 1 = 12 Z T (z)zdz: (2.6)


 (2L)3
L L
All the unknowns of the stress pro le in Equation (2.4) are now determined. The
mean strain "0 is always proportional to the average temperature in the plate while
the curvature arises from the asymmetry of the temperature distribution.
In the case of a symmetrical temperature pro le, T (z) = T ( z), the integral term
in Equation (2.6) is zero and the plate does not bend ( ! 1). The constant
temperature is a special case of a symmetrical thermal load. A non-constrained
plate with a constant temperature T expands freely and thus the resultant stresses
from (2.4) are zero:
T (z ) = T yields " = T and 1 = 0;
0

E  
 (z ) =
1  T + 0z T = 0: (2.7)
The other special case involves an unconstrained plate with the linear asymmetrical
temperature pro le:
T (z ) =
T z yields " = 0 and 1 = T ;
2L  0

 2L
E T T
 (z ) = 0 +
1  2L z
2L z = 0: (2.8)
The resultant stresses are again zero but this time the plate exhibits the resultant
curvature 1=. The multiplier T=2L in temperature pro le T (z) in Equation (2.8)
is the constant temperature gradient in the plate. The constant and the asym-
metrical temperature pro le can be combined together into the linear temperature
46
Chapter 2 2.1. Thermo-Elastic Analysis of a Flat Plate
undeformed state
T=T(z)

ρ
=_ T
ε0

ρ=∞ ε0

+ Tasym (z)

+
ε0 = 0
ρ

Trem (z)

ρ=∞ ε0 = 0

Figure 2.3: Decomposition of the temperature pro le T (z) into its constituent com-
ponents.
pro le, and as seen, this pro le does not contribute to thermal stresses. It is pos-
sible to extract these temperature pro les from an arbitrary temperature pro le
T (z ) as illustrated in Figure 2.3 and only the remaining temperature pro le Trem (z )
contributes to thermal stresses. The nal response of the material is given as the
sum of the constituent responses (superposition principle). For the symmetrical
temperature pro le, the stress distribution Equation (2.4) transforms into a simpler
form:
E  
 (z ) = T T (z ) : (2.9)
1 

47
Chapter 2 2.2. Simpli ed Viscoelastic Analysis
2.2 Simpli ed Viscoelastic Analysis
It was shown in Section 2.1 that it is not possible to induce residual stresses in
an elastic material by subjecting it to a thermal loading only. After the thermal
load is removed and material returns to a uniform temperature, the stresses drop
to zero. This is not how most polymeric materials behave: they show e ects of an
internally strained material at a uniform room temperature after thermal processing.
Thus, elastic constitutive model used in the analysis so far needs to be replaced with
the viscoelastic model which describes the behaviour of polymeric materials more
accurately. We shall focus our attention on a viscoelastic plate and a symmetrical
temperature distribution. From the Boltzmann superposition principle [14], the
equivalent of Equation (2.4) for a viscoelastic material is:
Z t
 (z; t) =
1 d ["
E (t  ) 0
T (z;  )]
d: (2.10)
1 0 d

In Equation (2.10), E (t  ) is the relaxation modulus and represents the response


of the material to a unit step in applied mechanical strain at time t =  (Figure
2.4).
E(t)
E(t-0) E(t-τ )

0 τ t

Figure 2.4: Relaxation modulus curves.


The stress pro le (z; t) is a sum of all relaxation processes, each of them initiated
by incremental mechanical strain applied at di erent times. The mean strain "0 can
48
Chapter 2 2.3. Simple Thermal Cycle
be eliminated from Equation (2.10) using the zero net-force boundary condition:
ZL Zt ZLZ t
d" @T (z;  )
 (z; t)dz = 0 ) E (t  ) 0 d = E (t  ) ddz:
d L @
0 0 0 0
(2.11)
By substituting this into Equation (2.10), the stress distribution becomes:
 L t Zt 
 (z; t) =
1 Z Z E (t  ) @T (z;  ) ddz E (t  )
@T (z;  )
d :
1  L
0 0
@
0
@
(2.12)
The former expression gives stresses in the viscoelastic plate that is subjected to a
symmetrical temperature distributions and is free of all external loads.

2.3 Simple Thermal Cycle


As an illustration of the use of Equation (2.12) we will subject the viscoelastic plate
to a simple thermal cycle. The cycle consists of imposing a known temperature
pro le, T (z), to the plate to at time t = 0. The plate is kept at T (z) until t = t0 ,
after which the temperature pro le is removed, Figure 2.5. This case is equivalent

T(z,t)
z
T(z)

z=L

t=0

t=t0
t

Figure 2.5: Simple thermal cycle.


49
Chapter 2 2.3. Simple Thermal Cycle
to applying a temperature increment T (z) at t = 0 and then applying the negative
temperature increment T (z) at t = t0 . The stresses remaining in the plate after
the cycle are the residual stresses since the material is at a uniform temperature.
The residual stress distribution in the plate after the simple thermal cycle is:
 ZL
1
res (z; t) =
1  L 0 E (t 0)T (z)dz E (t 0)T (z)
L 
1 Z E (t t0 )T (z )dz + E (t t0 )T (z ) ;
L
0
and (2.13)

Z L 
res (z; t) =
1
1  [E (t) E (t t0 )] L 0 T (z)dz T (z) : (2.14)
For the rst time we have managed to thermally induce residual stresses into the
material. When the material is exposed to T (z) for a long period of time (i.e. t0 !
1) the relaxation modulus E (t0 ) for the rst loading increment +T (z) converges to
zero, Figure 2.6, and the only remaining stresses are:
0)  ZL 
res (z; t ) =
E ( t 1
1  T (z) L 0 T (z)dz ; t0 = t t0 ; (2.15)
0

where t0 is the time after cooling down. There are many assumptions incorporated
into Equation (2.15), but it will be shown later, Equation (2.15) also represents the
response of the material for more general viscoelastic cases.
The situation where material relaxes completely during the thermal cycle is equiv-
alent to any process where fast stress relaxation process takes place. The rate of
stress relaxation is di erent at di erent temperatures: the higher the temperature,
the higher the rate of relaxation. Our current solution uses a single relaxation
function in the whole temperature range and does not account for the temperature
dependence in the relaxation modulus.

50
Chapter 2 2.4. Temperature Dependency of Viscoelastic Properties

E(t)

E(t) E(t-t0 )=E(t')

0 t0 t1 t2 t
z
0 t'

_ _

+ +

_ _
T(z) σres (z,t1 ) σres (z,t2 )

Figure 2.6: Residual stresses in the plate exposed to a long lasting simple cycle.
2.4 Temperature Dependency of Viscoelastic
Properties
In experiments, every stress relaxation curve is obtained isothermally at various
temperatures. The relaxation modulus curve for an arbitrary non-isothermal case
depends on the temperature history to which the material was subjected. It is de-
sirable to predict the relaxation modulus curve for an arbitrary non-isothermal case
from the relaxation curves recorded at constant temperatures. Such an interpolation
is the simplest for thermorheologically simple materials [19]. For thermorheologically
simple materials, all relaxation phenomena are accelerated with the rising tempera-
ture by a constant factor determined by experiment. The relaxation behaviour of a
thermorheologically simple material at any temperature can therefore be calculated
from a single relaxation curve at a reference temperature Tref , simply by stretching
or contracting the time scale for stress relaxation [15]. The new time scale d is
51
Chapter 2 2.4. Temperature Dependency of Viscoelastic Properties
related to the real time scale dt by:
d = (T )dt: (2.16)
The time{temperature transposing function, (T ), is equal to unity at the reference
temperature Tref . At Tref , the equivalent time scale d and the real time scale dt
coincide. For temperatures higher then the reference temperature, (T ) is greater
then 1, and the equivalent time, relevant for the stress relaxation, is increased. Each
point in the plate has its own equivalent time  , that is computed from the thermal
history as:
Zt
 (z; t) =  [T (z;  )] d: (2.17)
0

An increase in the temperature is equivalent to imposing material to changes for


a longer times,  instead of t, on the same stress relaxation curve (Figure 2.7a).
For thermorheologically simple materials, all relaxation curves have the same shape
when plotted on a log (time) abscissa. In such diagrams, the temperature change
simply causes a horizontal shift of the referenced relaxation curve for the value of
log (T ) (Figure 2.7b). The time{temperature transposing function (T ) is deter-
mined experimentally. There is a theoretical interpretation for the time{temperature
transposing function known as Williams-Landel-Ferry (WLF) equation [15, 17].
In order to include temperature e ects into the viscoelastic analysis, the relaxation
modulus function E (t) in Equation (2.10) is replaced by E ( ), giving the relation:
Z t
 (z; t) =
1 d ["0 T (z;  )]
1  0 E [ (z; t)  (z;  )]
d
d: (2.18)

Di erential/integral equations of the form (2.18) are mathematically complex and


cannot be solved analytically for most cases and require numerical solution. The
application of such a constitutive material model greatly depends on the accuracy
52
Chapter 2 2.5. Materials with Pronounced Softening Temperature
a) b)
E(t,T) E(t,T) T >Tref
Tref
E(t, Tref )=E( ξ )

log Φ( T )
Φ( Tref )=1

E(t,T) ; T >Tref

t1 t2 t
log (t)
0 ξ ξ2 ξ=Φ( T ) t ξ
1

Figure 2.7: Family of stress relaxation curves for a thermorheologically simple ma-
terial. a) Stress relaxation curve at temperature T is derived from the referenced
stress relaxation curve by stretching/contracting the time axis, b) horizontal shift
in the logaritmic diagram (b).
with which time{temperature transposing function (T ) is describing the real vis-
coelastic behaviour of the material.

2.5 Materials with Pronounced Softening Tem-


perature
Many polymeric materials show a tendency to become soft when they reach a certain
critical temperature Tg . The sti ness of the material is typically reduced by several
orders of magnitude above Tg due to very fast relaxation processes. In terms of the
time{temperature transposing function (T ), these materials can be described as
follows:
(T ) = 1 for T < Tg ;
(T ) = 1 for T  Tg ;
where temperature Tg is the softening temperature, also known as the glass tran-
sition temperature. The above equations characterise a material with unchanged
53
Chapter 2 2.5. Materials with Pronounced Softening Temperature
viscoelastic properties below Tg ) ( = t), while above Tg the equivalent time
for the stress relaxation tends to in nity ( ! 1), where the modulus is already
relaxed to zero. For materials with a distinct softening temperature Tg , thermal
stresses in the plate under a thermal load T (z; t), de ned by Equation (2.18), are
now expressed in one of the following two forms:
Z t
 (z; t) =
1 d ["0 T (z;  )]
1  0 E (t )
d
d for z < z < L

(2.19)
and
 (z; t) = 0 for 0 < z  z:
Here, z denotes the distance from the central plane to the solidi cation front. The
solidi cation front consists of all points on the isotherm Tg . The solidi cation front
divides the plate into two regions, Figure 2.8, and can be found from the expression
T (z ; t) = Tg . The region with the temperature above Tg is in the softened state where
thermal stresses are zero throughout. Below Tg the material is solid. Material that is
still in the softened state (T  Tg ) \wets" the previously solidi ed material surface
and after it cools down below Tg it becomes the part of solidi ed material core. The

Figure 2.8: Solidi cation front in a plate with the temperature pro le T (z; t).
54
Chapter 2 2.5. Materials with Pronounced Softening Temperature
thermal stress distribution must satisfy the zero net-force boundary condition if no
forces are applied to the plate at in nity. The softened region satis es this condition
automatically, while for the solid region the zero net-force condition is modi ed:
ZL
 (z; t)dz = 0: (2.20)
z
If we now subject the plate to a simple thermal cycle as shown in Figure 2.5, the
stress distribution for the region that was solid throughout the cycle, z < z < L, is:
 ZL
res (z; t) =
1 E (t)T (z )dz E (t)T (z )
1  L z
z
L 
1 Z E (t t0 )T (z )dz + E (t t0 )T (z ) ;
L
0

i.e.
 ZL
E (t)
res (z; t) = T (z )dz
1  L z
z
ZL 
E (t t0 ) h i
T (z )dz E (t) E (t t0 ) T (z ) : (2.21)
L
0

Similarly, for the region softened in the cycle, 0 < z  z, we have:
 Zz
res (z; t) =
1 E (1)T (z)dz E (1)T (z )
1  z
0
L 
1 Z E (t t0 )T (z )dz + E (t t0 )T (z ) ;
L
0

i.e.
Z  L 
res (z; t) =
1
1  E (t t0 ) L 0 T (z)dz T (z ) : (2.22)

55
Chapter 2 2.6. Solution for Series of Simple Thermal Cycles
If the material is exposed to a simple thermal cycle for a short period of time, i.e.
t0  t t0 , the di erences in relaxation modulus E (t) and E (t t0 ) are negligible.
Taking this into account, the nal expressions for the residual stress pro les are:
  ZL ZL 
E (t0 ) 1 1
res (z; t0 )  T (z )dz T (z )dz z < z < L;
1  L 0
L z
z
(2.23a)
  ZL 
E (t0 ) 1
res (z; t0 )  T (z )dz T (z ) 0 < z < z:
1  L 0
(2.23b)
Equations (2.23a) and (2.23b) have a form which is identical to the Hooke's law
(2.4) where the bracketed term represents the strain in the plate. Therefore, we
can consider the bracketed terms again to be the strains, \frozen-in" the material
during the short thermal cycle. Frozen-in strains contain the relevant information
of the thermal history to which the material has been subjected. They are constant
in the region which remained solid in the cycle, while in the region softened in the
cycle, there is a distribution of strain. This solution is of theoretical importance
only because it is impossible in practice to simulate instantaneous heating/cooling
processes as required for a simple thermal cycle.

2.6 Solution for Series of Simple Thermal Cycles


If the plate is subjected to another simple thermal cycle soon after the previous
one, residual stresses will be in uenced only in the newly solidi ed region. In the
rest of the plate, the distribution of residual stresses from previous thermal cycles
remains preserved: it can only have been increased or decreased by a constant. The
continuous cooling process from a fully softened condition, which describes practical
conditions, can be modelled as a series of simple cycles, Figure 2.9. During the
56
Chapter 2 2.6. Solution for Series of Simple Thermal Cycles

T(z,t)

T=Tg

z t

T(z,t)
z=L

T=Tg

z t

z=L

Figure 2.9: Simulation of the continuous cooling process as a series of simple cycles.
continuous cooling of an initially softened plate, the solidi cation front moves from
the outer surfaces towards the inner sections of the plate. In the cooling process,
the point on the current solidi cation front z = z(t) will end in the solid region at
time t + dt. While transforming from a softened into a solid like state, the strain at
the interface will become frozen in the material. After solidi cation no additional
strain is induced. The residual stress gradient is therefore given as a gradient on the
residual stress distribution at the interface z = z, after a short thermal cycle. By

57
Chapter 2 2.7. Analogy Between Thermal and Residual Stresses
di erentiating the residual stress distribution (2.23b) we obtain:

@res (z; t0 ) E (t) @T (z; t0 )
@z z=z 1 
=
@z z=z
: (2.24)
The nal residual stress pro le after continuous cooling is obtained by integration
of all frozen-in stress increments to give:
 Zz 
E (t) @T (z 0 ; t0 )
res (z; t ) =
1  0 @z0 z =z dz + "0 : (2.25)
0 0
0

The integration constant "0 is such to satisfy zero net-force boundary condition and
is equal to:
Z L  Zz 
"0 =
1
@T (z 0 ; t0 )
dz 0 dz: (2.26)
L @z 0 z =z 0
0 0

Equation (2.25) predicts the residual stresses in a viscoelastic plate cooled symmet-
rically and consists of two terms. The rst term comprises the elastic constants of
the material, the combination of which depends on the assumed stress state (plane
stress, plane strain, rotational symmetry). The second term describes the frozen-
in strains accumulated in the material during the cooling process and contains all
relevant information from the thermal history. The gradient on the frozen-in strain
distribution is proportional to the temperature gradient at the moment of solidi ca-
tion, which is how the thermal history and mechanical response are coupled together
for this case.

2.7 Analogy Between Thermal and Residual Stresses


A further analogy to the thermal stress problem is possible by the following substi-
tution:
Zz Z0
@T (z 0 ; t0 ) @T (z 0 ; t0 )
T (z ) =
res dz =
0 dz 0 : (2.27)
@z 0
z =z
0 @z 0
z =z
0
0 z

58
Chapter 2 2.7. Analogy Between Thermal and Residual Stresses
The new auxiliary function Tres(z) is called the \residual" temperature eld. The
average value of Tres through the thickness of the plate is T res. After a long time
spent at the room temperature, the relaxation modulus E (t0) of the material from
Equation (2.25) becomes equal to the long term modulus at room temperature E1.
The residual stress distribution, from Equation 2.25, now becomes:
E1  
res (z ) = Tres (z ) : (2.28)
1  T res
The form of Equation (2.28) is identical to the solution of thermal stresses in a linear
elastic material, described by Equation (2.9). Therefore, the thermo-mechanical
response of the material to an applied residual temperature eld Tres is identical to
the response to frozen-in strains. By knowing the residual temperature eld Tres, the
residual stress problem can be considered as a thermal stress problem, that can be
solved using conventional numerical methods. The key point in the residual stress
analysis remains the determination of the residual temperature eld.

59
Chapter 3
Temperature Solutions for a
Convectively Cooled Plate

Residual stresses in a polymeric plate are clearly linked with the thermal history
during thermal processing, which are not known in advance. In this Chapter, some
analytical solutions for temperature distribution in a at plate geometry subjected
to various initial and boundary conditions are brie y summarised.

3.1 Determination of Temperature Fields


In practical applications temperature elds can be established in several ways: ex-
perimentally, analytically and numerically. In experimental methods, temperatures
are measured at a number of locations on and within the plate. The complete
temperature map required for residual stress calculations can subsequently be ap-
proximated from the captured discrete data. This mapping procedure can turn into
a diÆcult and error-prone operation and can introduce a number of uncertainties
into the temperature history. The number of measured temperature signals is usu-
ally limited. Also, the actual temperature distribution between the measured points
can di er signi cantly from the assumed temperature distributions, typically lin-
ear or parabolic. There exist a small number of analytical solutions available for
60
Chapter 3 3.1. Determination of Temperature Fields
the temperature T . These are often mathematically very complex expressions and
are only applicable to a limited number of situations, usually involving idealised
geometries and idealised material properties. For realistic situations, approximate
numerical solutions are the only viable options. However, analytical solutions are
still very important as they can be used to validate the numerical solutions.

3.1.1 Equation of the Heat Conduction


Numerical methods are widely employed for determining temperature elds. In
numerical methods the behaviour of the thermal system is described mathematically.
The mathematical description includes all parameters of the thermal system such
as material behaviour, initial thermal state and thermal interactions of the material
with the surrounding media, either solid or uid. Temperature histories are now
solutions of the conventional thermal energy conservation equation [57]:
Z Z Z
d
dt
cT dV = n  q dS + hV dV: (3.1)
V S V
Here V is the volume of an arbitrary part of the body bounded by the surface S
with the outward pointing unit normal vector n, T is the temperature,  is the
density, c is the speci c heat capacity, q is the heat ux vector and hV is the
intensity of the volumetric heat source. The Fourier law of heat conduction describes
the relationship between the heat ux and the temperature variation in the body.
Mathematically it is expressed as:
q = krT; (3.2)
where k is the thermal conductivity of the material and rT is the temperature
gradient vector. The energy conservation equation (3.1) now becomes:
Z Z Z
d
dt
cT dV = n  (krT ) dS + hV dV: (3.3)
V S V

61
Chapter 3 3.1. Determination of Temperature Fields
Equation (3.3) is the integral form of the energy conservation law expressed with
the single unknown variable T . The unique solution for temperature T = T (r; t)
can be obtained for any material point r at any time instant t by specifying the
initial temperature conditions of the body and temperature conditions on the body
boundaries. The solution T (r; t) represents the thermal history that is required for
residual stress calculations.
It is common in continuum mechanics to express the energy conservation equation in
the alternative di erential form. With the use of Gauss divergence theorem [58], the
surface integral term in Equation (3.3) can be transformed into a volume integral
so we have:
Z Z Z
d
dt
cT dV = r (krT )dV + hV dV: (3.4)
V V V
Since volume V is arbitrary, it follows:
@ (cT )
@t
= r (krT )+ h : V (3.5)
Equation (3.5) is the di erential form of the Equation (3.4) for the solid material.
In the special case with the constant material properties (; c; k) and by adopting
 = k=(c), the energy equation (3.5) is simpli ed into:
@T
@t
= r2 T: (3.6)
This equation is commonly known as the equation of heat conduction. The newly
de ned material property  is the thermal di usivity. Physically, the higher the
value of  the faster the system will stabilise thermally. Also note that in the last
equation the volumetric heat source/sink term hV is neglected. In the case in which
T does not vary with the time, Equation (3.6) reduces to the Laplace's equation:
r2 T = @@xT2 + @@yT2 + @@zT2 = 0:
2 2 2
(3.7)

62
Chapter 3 3.2. One-Dimensional Analytical Temperature Solution
3.1.2 Thermal Boundary Conditions
Temperature conditions on the boundaries of the solid can be given in many di erent
forms. In the engineering practice we can often measure the temperature information
either in terms of its value or as a temperature gradient. The latter covers many
engineering appliances, i.e. cases where the heat is supplied to a solid by a heater,
cases with insulated surfaces or cases where the surfaces of the solid are in contact
with uid (convective cooling). Convective cooling conditions are important for
describing cooling processes in polymeric plate manufacture, as the heat from hot
plate surfaces is conducted to a cooling medium. The empirical Newton's law for
convection quanti es the amount of heat per unit area qb conducted from a solid
surface into the surrounding uid as [59]:
qb = h(Tb T1 ); (3.8)
where Tb is the temperature of the solid boundary and T1 is the temperature of
the surrounding uid. The factor of proportionality h in Equation (3.8) is the
local heat transfer coeÆcient. The heat transfer coeÆcient is an integral quantity
which quanti es, in an average sense, all heat transfer processes within the thermal
boundary layer of the uid next to a solid surface. In general, it depends on the
uid ow eld, shape and size of the solid body, material properties of the uid, etc.
There are various correlations in the literature that quantify h for di erent cooling
systems.

3.2 One-Dimensional Analytical Temperature So-


lution
Let us now examine various analytical solutions relevant to cases of convective cool-
ing of a at plate. The derivation of these solutions will be based on a slab solutions
63
Chapter 3 3.2. One-Dimensional Analytical Temperature Solution
presented by Carslaw and Jaeger [24]. Consider a solid plate bounded by a pair
of parallel planes, z = L and z = L. Temperature variation exists only in the
thickness direction z. On both boundary surfaces the heat is conducted into the
surrounding uid at zero temperature T1 = Tc = 0ÆC. Convective cooling condi-
tions will be assumed identical on both sides of the plate. The initial temperature
of the plate is T0 (z). This thermal system is mathematically described with the
one-dimensional heat conduction equation:
@T
=  @@zT2
2

@t
for L  z  L: (3.9)
From the theory of partial di erential equations, it can be shown that the following
expression satis es the heat conduction Equation (3.9):
T (z; t) = Ae  2 t (C cos z + D sin z ); (3.10)
where A, C , D and are arbitrary constants. The temperature in the plate T (z; t)
is uniquely de ned only after specifying the initial temperature distribution in the
plate and the boundary conditions. On boundary surfaces, z = L and z = L, the
heat conducted from the plate is passed entirely onto the cooling medium. Thus,
we can equate the Fourier's equation (3.2) and the Newton's law of cooling (3.8) to
obtain two boundary conditions:
@T
k
@z
+ hT = 0 at z = L;
@T (3.11)
k
@z
+ hT = 0 at z = L:

The temperature solution in the solid (3.10) must also satisfy these conditions.
Substitution of T and its spatial derivative @T=@z from (3.10) into (3.11) yields the
following two conditions:
C (H cos L sin L) D (H sin L + cos L) = 0;
(3.12)
C (H cos L sin L) + D (H sin L + cos L) = 0;
64
Chapter 3 3.2. One-Dimensional Analytical Temperature Solution
where h=k = H . The non-zero solution for C and D exists if the determinant of the
system (3.12) is zero [60]. From this we obtain the following condition:
(H cos L sin L) (H sin L + cos L) = 0; (3.13)
or
cot (2 L) = 2 H 2H : (3.14)
This is the eigen-condition for the system. This equation has multiple solutions
when solving for due to the periodical nature of the cotangent function cot (2 L).
The eigen condition is mathematically complex and solutions n, where n is the
n-th positive root of the Equation (3.14), cannot be expressed explicitly. Figure 3.1
illustrates the distribution of roots n.

Figure 3.1: The distribution of the roots n of the eigen-equation.


With the eigen condition satis ed it is clear that with the substitutions:
C = (H sin L + cos L) ; (3.15)
D = (H cos L sin L) ; (3.16)
65
Chapter 3 3.2. One-Dimensional Analytical Temperature Solution
both conditions from Equation (3.12) are automatically satis ed. The newly calcu-
lated pairs of constants C and D are constants for the temperature solution (3.10).
It is useful to note that C 2 + D2 = 2 + H 2. Due to multiple roots of the eigen-
equation = n, there are multiple pairs of C = Cn and D = Dn that satisfy the
boundary conditions equation. Each pair of constants Cn and Dn can now be used
for the temperature solution equation:
T (z; t) = An (Cn cos n z + Dn sin n z )e  2n t : (3.17)
In the temperature solution (3.17) we still have not incorporated the initial temper-
ature condition. Rewriting Equation (3.17) for t = 0, it follows:
T (z; 0) = An (Cn cos n z + Dn sin n z ) ; (3.18)
from which one can see see that Equation (3.17) is the temperature solution only if
the initial temperature is T0 (z) = An(Cn cos nz + Dn sin nz ). Such a special case
is almost impossible in the engineering practice, rendering Equation (3.17) only of
limited importance. However, solution (3.17) can be considered as the simplest
transient temperature solution for a plate subjected to convective cooling. It can be
used as an approximate solution or for validation purposes of numerical solutions.
To account for any realistic initial temperature pro le, we can assume that T0 (z)
can be expressed as an in nite series:
1
X
T0 (z ) = A1 Z1 + A2 Z2 + ::: + An Zn + ::: = An Zn ; (3.19)
n=1

where
Zn = Cn cos n z + Dn sin n z: (3.20)
The substitution for Zn, Equation 3.20, is introduced here for convenience. The
series (3.19) is a linear a combination of terms Zn, each term being formed from Cn
66
Chapter 3 3.2. One-Dimensional Analytical Temperature Solution
and Dn which satisfy the eigen condition, Equation (3.14). This is to ensure that
boundary conditions are satis ed throughout. The unknown coeÆcients An may be
obtained in the form similar to Fourier series. From the weighted integration of the
series (3.19) we get:
ZL ZL
T0 (z )Zm dz = (A1 Z1 + A2 Z2 + ::: + AnZn) Zmdz: (3.21)
L L

The right hand-side integral can be evaluated further by noticing that:


ZL ZL

Zm Zn dz = 0 (m 6= n) and ZnZn dz = n2 + H 2 L+H: (3.22)
L L

These two conditions leave a single integral term on the right hand side of Equa-
tion (3.21) from which the unknown constants of the series An are now uniquely
determined:
ZL ZL
 
T0 (z )Zn dz = An ZnZn dz = An ( n2 + H 2 )L + H ) ; (3.23)
L L
or
ZL
An =
1 T0 (z )Zn dz: (3.24)
( n2 + H 2)L + H L

This brings us to the end of the derivation for the transient temperature distribution
in the plate:
1
X C cos z + D sin n z
T (z; t) = e  t n 2 n 2 n
2

( n + H )L + H
n

n=1
ZL
 T0(z)(Cn cos nz + Dn sin nz )dz (3.25)
L

The former equation is valid for determining temperature pro les in the a plate of a
thickness 2L, cooled from the initial temperature pro le T0(z) by applying a cooling
67
Chapter 3 3.2. One-Dimensional Analytical Temperature Solution
medium of a zero temperature Tc = 0Æ C on both boundary surfaces. If the cooling
medium is at di erent temperature then 0ÆC, temperature terms in (3.25) can be
considered as temperature di erences to the cooling temperature Tc:
1
X C cos z + D sin n z
T (z; t) Tc = e  t n 2 n 2 n
2

( n + H )L + H
n

n=1
ZL
 [T0(z) Tc](Cn cos nz + Dn sin nz)dz; (3.26)
L

from which T (z; t) is easily determined. In the special case of the symmetrical initial
temperature pro le T0 (z), the expression (3.26) is simpli ed:
1 ZL
X Cn2 cos n z
T (z; t) = Tc + e  2n t [T0 (z) Tc ] cos n zdz: (3.27)
n=1
( n2 + H 2)L + H L

This equation gives a fully symmetrical solution for temperatures in the plate. In the
symmetrical case there is no heat ow across the central plane of the plate (z = 0).
Therefore, Equation (3.27) is equally applicable for cooling cases with an adiabatic
surface at z = 0. This can be used for simulating well insulated surfaces. In many
engineering appliances the cooling commences from an uniform initial temperature
T0 . This simpli es the temperature solution further into:
1
X Cn2 cos n z 2 sin nL :
T (z; t) = Tc + e  t ( ) (3.28)
2
T T
( n2 + H 2)L + H
n
0 c
n=1
n

68
Chapter 3 3.3. Illustration of the Analytical Solution
3.3 Illustration of the Analytical Solution
To illustrate the use of the derived equations, we will now consider two cases of one-
dimensional heat transfer in a plate: symmetrical and asymmetrical cooling case.
In both cases, the thickness of the plate is 10 mm and the initial temperature is
uniform at T0 = 200ÆC.

3.3.1 Symmetrical Temperature Solution


In the symmetrical cooling case convective cooling commences from both boundary
surfaces at time t = 0. The cooling medium on both sides is at zero temperature
and the heat transfer rate is described with the constant heat transfer coeÆcient h.
The material properties and the cooling parameters are given in Table 3.1.
 (kg/m3 ) k (W/mK) c (J/kgK)  (m2 /s)
960 0.4 2000 2:083  10 7

L (mm) h (W/m2 K) H (1/m) T0 (Æ C) Tc (Æ C)


5 600 1500 200 0

Table 3.1: Material properties and cooling parameters for the symmetrical cooling.
The roots n of the eigen-equation (3.14) can be evaluated numerically. As seen
in Figure 3.1, the roots n are regularly displaced and there is one root n in
successive =2L intervals. This helps to determine a good initial guess for each
root and the solutions of n can be obtained in a few Newton-Raphson cycles.
The eigen-equation was solved using the Mathcad software package [61], and the
implementation procedure is given in Figure 3.2. The only input to the procedure
is the number of required roots n. The more roots are taken into consideration, the
better the series, Equation (3.19), describes the initial temperature distribution, in
our case constant 200ÆC. Throughout this analysis 100 roots have been considered,
and this solution will be referred as the \full" solution. The rst 10 roots n and
69
Chapter 3 3.3. Illustration of the Analytical Solution

Mathcad implementation procedure:


x H
F_left( x) cot( 2 . x. L ) F_right( x )
2. H 2. x
1. π
roots_α( N ) α guess
2 2. L
for n ∈ 0 .. N 1
π
α guess α guess
2. L
αn root F_left α guess F_right α guess , α guess

α guess αn
1
α. m

277.565
1
roots_α( 3 ) = 557.186
m
840.345

Figure 3.2: Mathcad implementation procedure for determining roots n of the


eigen-equation for the convective cooling temperature solution.

n Cn Dn
n
m 1

1 277.565 1525.465 0
2 557.186 0 -1600.143
3 840.345 -1719.355 0
4 1127.706 0 1876.625
5 1419.248 2065.009 0
6 1714.547 0 -2278.085
7 2013.033 -2510.438 0
8 2314.135 0 2757.757
9 2617.355 3016.711 0
10 2922.279 0 -3284.770

Table 3.2: Auxiliary constants for the temperature solution equation. The numerical
values are speci c for the cooling system described in Table 3.1.

70
Chapter 3 3.3. Illustration of the Analytical Solution
the corresponding values for auxiliary constants Cn and Dn, for the conditions listed
in Table 3.1, are presented in Table 3.2.
In the temperature calculation, the temperatures are monitored at ve points across
the plate half-thickness with one in the middle, one on the top surface and others
equally displaced between them. The temperature histories in these points are
shown in Figure 3.3. In the early stages of cooling, the temperatures on the cooled
200
z=5.00 mm (top surface)
z=3.75 mm
z=2.50 mm
150
z=1.25 mm
temperature ( o C)

z=0.00 mm (middle surface)

100

50

0 60 120 180 240


cooling time (s)
5.0
thickness coordinate z (mm)

1 0
5 2
10
2.5 20
40 30
60 50
120 90
0.0

-2.5

-5.0
0 50 100 150 200
temperature ( o C)

Figure 3.3: Temperature histories in the symmetrically cooled plate from the initial
uniform temperature.

71
Chapter 3 3.3. Illustration of the Analytical Solution
surfaces drop very rapidly. It will take some time until the temperature in the inner
region of the plate starts dropping, due to initially low, but existing temperature
gradients. The established temperature di erences between the cooling surfaces and
the interior are a function of the thermal di usivity , and consequently all other
properties de ning thermal difusivity. Materials with the higher heat capacity c
accumulate more thermal energy that must be conducted from the material during
cooling. This will increase the temperature di erences during cooling. Opposite to
it, materials with higher thermal conductivity k will remove the accumulated heat
quicker, lowering the temperature di erences in the plate. For polymeric materials
the thermal difusivity is typically as low as  ' 10 7 m2/s [62] and hence the
temperature di erences during cooling will be high. In Figure 3.3 the temperature
pro les at speci ed times are presented. At longer cooling times these pro les turn
into regular parabolic curves with varying amplitude. At the nal stages of cooling,
the temperature pro les are converging towards the uniform temperature of the
cooling medium Tc.

3.3.2 Asymmetrical Temperature Solution


In the asymmetrical case the convective cooling applied was only to the top surface,
while the bottom surface was kept perfectly insulated. This is equivalent to case
of symmetrical cooling of plate with the thickness of 20 mm (L = 10 mm). The
corresponding temperature histories and temperature distribution graphs are given
in Figure 3.4. As expected, temperature gradients for all temperature pro les are
zero on the bottom surface.
The last two examples show typical temperature history traces when convective
cooling is employed. Qualitatively, any realistic cooling case with non-constant heat
transfer coeÆcient h and temperature-dependent material properties will exhibit
similar behaviour.
72
Chapter 3 3.4. \ln-cos" Residual Temperature Solution

200
z=10.00 mm (top surface)
z=7.50 mm
z=5.00 mm
150
z=2.50 mm
temperature ( o C)

z=0.00 mm (bottom surface)

100

50

0 120 240 360 480 600


cooling time (s)
10.0
thickness coordinate z (mm)

0
2 1
5
10
7.5 30 20
60
120
5.0 180
240
300
360
2.5

0.0
0 50 100 150 200
temperature ( o C)

Figure 3.4: Temperature histories in the asymmetrically cooled plate.


3.4 \ln-cos" Residual Temperature Solution
The mechanical response of the material to cooling conditions is of a prime interest.
It was identi ed previously that for materials with a softening temperature Tg , the
temperature gradient at the moment of solidi cation is a measure of the frozen-in
strains. The temperature gradients at the moment of solidi cation are integrated
(Equation 2.27) to form the residual temperature eld Tres. This section describes
the simpli ed analytical solution for Tres derived from the temperature solution for
73
Chapter 3 3.4. \ln-cos" Residual Temperature Solution
the convective cooling case.
For the cases considered, the temperature gradient for any point z and at any time
t can be obtained from the temperature distribution (3.26) by di erentiating it with
the respect of z. Our aim is to obtain temperature gradient at the moment of
solidi cation for any point z = z, where T = Tg . This requires the appropriate time
t = tsol to be substituted into the temperature solution. The tsol is the solidi cation
time function and can be deducted from the condition:
T (z ; tsol ) = Tg ; (3.29)
thus giving
tsol = tsol (z ): (3.30)
Although this condition gives the unique solution for the solidi cation time tsol , it
cannot be expressed explicitly from the in nite series temperature solution (3.26).
The complexity in solving for tsol does not allow many analytical solutions for Tres(z)
to be derived analytically. However an analytical solution for Tres(z) exists when a
single series term is used to describe the temperature history. If only the rst term
in the series (3.26) is chosen, i.e. (n = 1) ) (C1 6= 0; D1 = 0), it follows:
ZL
C12 cos 1 z
T (z; t) = Tc +e  21 t [T0 (z) Tc ] cos 1 zdz: (3.31)
( 12 + H 2)L + H L
This is the dominant term in the series expansion at longer cooling times t. The
equivalent simpli ed form of Equation (3.31) is:
T (z; t) = Tc + F1 (t)F2 (z ): (3.32)
This is the temperature pro le of a xed shape F2 (z) scaled in time with F1(t).
Williams [31] used temperature pro les of the form (3.32) to analytically derive the
residual temperature eld and will be outlined below.
74
Chapter 3 3.4. \ln-cos" Residual Temperature Solution
The temperature pro le at the time when point z = z reached solidi cation tem-
perature Tg is:
T [z; tsol (z )] = Tc + F1 [tsol (z )]F2 (z ); (3.33)
or, speci cally, for the solidi cation point z:
T [z ; tsol (z )] = Tc + F1 [tsol (z )]F2 (z ) = Tg : (3.34)
By eliminating F1[tsol (z )] we obtain:
F (z )
T [z; tsol (z )] = Tc +(Tg Tc ) 2 : (3.35)
F2 (z )
The temperature T is no longer an explicit function of the solidi cation time tsol .
The solidi cation time information is indirectly supplied into the analysis through
the solidi cation front location variable z. The temperature gradient is obtained by
di erentiating Equation (3.35):
@T [z; tsol (z )] Tg Tc dF2 (z )
@z
= F (z)  dz (3.36)
2

and in the solidi cation point z = z it becomes:



@T [z; tsol (z )] Tg Tc dF2 (z )
@z = F (z )  dz : (3.37)
z =z 2 z =z
The integration of temperature gradients in (3.37) results the residual temperature
eld Tres:
Zz Zz Zz
dF2 (z )
0 0

dTres (z ) = dT [z ; tsol (z )] = (Tg Tc ) ; (3.38)


F2 (z )
z
0 z z
or
F (z )
Tres (z ) = Tres(z0 ) (Tg Tc ) ln 2 : (3.39)
F2 (z0 )
The integration limits in the 2nd integral in (3.38) are reversed deliberately to pro-
duce positive values for Tres. In any cooling case there is a hotter region and a point
75
Chapter 3 3.4. \ln-cos" Residual Temperature Solution
in there z0 that will reach the solidi cation temperature Tg last. The positive tem-
perature gradients are always pointing towards point z0 during cooling. If the upper
integration limit is set to this point, that would ensure positive Tres throughout.
The integration constant Tres(z0 ) is chosen zero in the point which solidi es last.
Using the temperature solution (3.31) and noticing that the point that solidi es
last is in the middle of the plate at z0 = 0, a simple expression for the residual
temperature eld is derived:
Tres (z ) = (Tg Tc ) lncos 1 z: (3.40)
From here, the residual temperatures are increasing with the lower cooling temper-
ature Tc and with higher the rst root 1 (more intensive cooling h ", lower thermal
conductivity k #). The \ln-cos" form of the residual temperature solution was used
in the past by Bartenev [29].

3.4.1 Illustration of the \ln-cos" Solution


The residual temperature solution will be illustrated on two cooling examples, again
the symmetrical and the asymmetrical cooling. The heat transfer coeÆcient of
h = 100 W/m2K is assumed for all cooling surfaces. The material properties
remain as those in Table 3.1. Unlike previous cases, here we shall adopt the solution
with a single term in the series expansion (3.31). The term T0 (z) represents the
initial temperature distribution imposed on the system only if the full solution is
considered. With the single term in the solution, the temperature distribution at
zero time T (z; 0) is just the rst approximation of the cooling case with the initial
temperature T0 (z). For the constant initial temperature T0 (z) = T0 the solution
(3.31) becomes:
C 2 cos z 2 sin 1L :
T (z; t) = Tc + e  t 2 1 2 1 ( T0 Tc ) (3.41)
2

( 1 + H )L + H
1
1

76
Chapter 3 3.4. \ln-cos" Residual Temperature Solution

z=5.00 mm (top surface)


200 z=3.75 mm
z=2.50 mm
z=1.25 mm
temperature ( o C)

150 z=0.00 mm (middle surface)

Tg
100

50 42.5 s
113.9 s

0 60 120 180 240


cooling time (s)
5.0
thickness coordinate z (mm)

113.9 42.5

2.5 Tg
− Tres (z )
0.0
0
20
-2.5 60
100 80
140

-5.0
-50 0 50 100 150 200
o
temperature ( C)

Figure 3.5: Temperature histories and the residual temperature distribution for
symmetrically cooled plate (2L = 10 mm, h = 100 W/m2K, 1 = 186:151 m 1,
C1 = 311:693 m 1 ).

The solution of Equation (3.31) for the symmetrical cooling case (2L = 10 mm,
T0 = 200Æ C, Tc = 0Æ C) is shown in Figure 3.5. All temperatures in the plate are
initially above the assumed solidi cation temperature Tg = 100ÆC. The temperature
on the cooled surfaces will reach Tg at 42.5 s from the start of cooling. This is
when the solidi cation process initiates at both plate surfaces. From this moment
the frozen-in strains start to accumulate in the plate. The solidi cation front is

77
Chapter 3 3.4. \ln-cos" Residual Temperature Solution
gradually moving towards the interior of the plate. At time 113.9 s the front reaches
the middle surface and solidi cation is complete. The residual temperature eld is
formed during these two events.
The slope of the temperature pro le during solidi cation is preserved in the residual
temperature eld and is the measure of frozen-in strains. In Figure 3.5 the calculated
residual temperature eld from the \ln-cos" solution is shown. The initial slopes
on it are shown with dashed lines. These slopes match slopes on the temperature
pro le when solidi cation started. All other slopes on Tres match the slopes on the
temperature distributions at the moment of solidi cation. The residual temperature
value is zero in the middle point of the plate. The residual temperature gradient in
this point is also zero.
The solution (3.31) for the asymmetrical cooling case (L = 10 mm, T0 = 200ÆC,
Tc = 0Æ C) is presented in Figure 3.6. The heat is convected from the top surface
only, while the bottom surface is perfectly insulated. The initial temperature on the
top surface of the plate is below the assumed solidi cation temperature Tg = 100ÆC.
The solidi cation data is therefore unknown for a few points close to the top surface
which are initially solid. The residual temperature solution (3.40) would still produce
the solution since it is mathematically valid for all points, even those that reached
solidi cation temperature at \negative" times. However, for a proper illustration
of the residual temperature concept we will overcome this problem by decreasing
the solidi cation temperature to Tg = 90ÆC. This way the whole plate is initially
softened and a complete solidi cation picture can be obtained during cooling. The
top surface will now reach 90ÆC after 36.7 seconds while the last point on the
insulated bottom surface will solidify after 359.8 seconds. Yet again, the residual
temperature is highest on the cooled boundary and is set to zero at the point that
solidi ed last. The slopes from the cooling temperature pro les at solidi cation are

78
Chapter 3 3.4. \ln-cos" Residual Temperature Solution
250

z=10.00 mm (top surface)


200 z=7.50 mm
z=5.00 mm
z=2.50 mm
temperature ( o C)

150 z=0.00 mm (bottom surface)

100 Tg

50 36.7 s
359.8 s

0 120 240 360 480 600 720 840


cooling time (s)
10.0
thickness coordinate z (mm)

7.5 359.8 36.7

Tg
5.0
− Tres (z )
0
60
120
2.5 180
240
300
420
0.0
-100 -50 0 50 100 150 200 250
temperature ( o C)

Figure 3.6: Temperature histories and the residual temperature distribution for
asymmetrically cooled plate (L = 10 mm, h = 100 W/m2K, 1 = 114:223 m 1,
C1 = 274:858 m 1 ).

identical to the slopes on the residual temperature. The two former cases will be
used later to validate the numerical procedures for obtaining Tres.

3.4.2 Application of the \ln-cos" Solution


It still remains to discuss the applicability of the \ln-cos" solution for a realistic
cooling cases. The \ln-cos" residual temperature solution was derived from the
79
Chapter 3 3.4. \ln-cos" Residual Temperature Solution
temperature solution which is only a rst approximation of the general temperature
solution for convective cooling cases. This rst approximation does not describes
initial temperature pro le in the plate well, nor does it capture the temperatures in
the plate at times shortly after cooling commenced. At later times, the situation
becomes more acceptable. The exponential term in the general temperature solution
e  t reduces as the root n increases. At longer times, the main contribution to
2
n

the temperature solution will arise from the leading root 1 . If the material starts
solidifying at times when the temperature solution is represented well with the
simpli ed solution, the presented residual temperature solution should be acceptable
for residual stress analysis.
In Figure 3.7, a full and simpli ed temperature solution for an arbitrary symmet-
rical cooling case is presented. The number of terms in the series solution used is
n = 1 and n = 100 respectively. As expected, both solutions di er only in the early

full temperature solution (n=100)


200 aproximate solution (n=1)

middle surface
temperature ( o C)

150 Tg1

Tg 2
100
Tg 3

50

top/bottom surface
0

0 60 120 180 240 300


cooling time (s)

Figure 3.7: Comparison of the full and simpli ed temperature solution for an ar-
bitrary symmetrical convective cooling case (2L = 10 mm, h = 100 W/m2K,
1 = 186:151 m 1 , C1 = 311:693 m 1 ).

80
Chapter 3 3.4. \ln-cos" Residual Temperature Solution
stages of cooling. If the solidi cation temperature is high (Tg = Tg1 in Figure 3.7),
solidi cation will start very early in the cooling process. In the initial time interval
the simpli ed temperature solution does not match the full solution well. Still, the
\ln-cos" residual temperature solution gives a good indication of the actual resid-
ual temperature eld. It would, however, produce a conservative estimate of the
solution, since the temperature gradients are higher in the real case. If the solidi -
cation starts at later times (Tg = Tg2 or Tg3 in Figure 3.7), the di erences between
two temperature solutions are less obvious and the \ln-cos" residual temperature
solution produces a very accurate solution.

81
Chapter 4
The Residual Temperature
Concept

The solution for the residual stresses in a viscoelastic plate has been derived in
Chapter 2. This solution is applicable for cases of one-dimensional symmetrical
solidi cation and is suited for materials with well-pronounced softening tempera-
ture Tg . This Chapter expands the proposed residual temperature analysis to more
general solidi cation cases and multidimensional geometries.

4.1 Frozen-in Strains


It has been shown that during solidi cation the strains are frozen-in the polymeric
material. As the solidi cation continues, there exists a distinct interface between
the softened and solid material (Figure 2.8), i.e. the solidi cation front. Two
neighboring layers of the material on di erent sides of the interface behave di erently.
On the softened side, the material cannot sustain any shear force. The material
\ ows" and the stresses relax. The \ ow" here is a general term for all relaxation
processes that material might experience. Some polymeric materials behave like
non-Newtonian or Bingham uids and they ow in a uid-like manner when above
Tg . Others are in a rubbery state, where polymer molecules can rearrange into

82
Chapter 4 4.1. Frozen-in Strains
lower stress states [63]. The material on the solid side of the interface can sustain
mechanical strains, and stresses can be generated. During solidi cation the softened
material \sticks" to the solidi ed material core. There is a di erential temperature
between layers on opposite sides of the interface T , and the layers are expanded
accordingly via the coeÆcient of the linear thermal expansion. After the new time
increment t these layers will nd themselves in the solid region. The volumetric
increment of frozen-in strain "fr has just been generated and stresses the material
from that moment onwards. At the end of cooling, when all temperature gradients
are removed, the strain remains \locked" in the material. The frozen-in strain can
be expressed as:
"fr = (Tg )T: (4.1)
The factor of proportionality between "fr and temperature increment T in (4.1)
is the coeÆcient of linear thermoexpansion . The referenced temperature for is
the softening temperature Tg because all strain increments become frozen-in at this
temperature. Let us consider the Taylor series expansion in space of a temperature
pro le around a newly solidi ed point r [58]. With only the rst term in the series,
T becomes:

@T
T = @r  r: (4.2)
T =Tg

Here "  " is the inner or dot product operator between two vectors, creating a scalar.
Substituting T into Equation (4.1) we have:
" #
@T
"fr = (Tg ) @r T =Tg
 r : (4.3)
When an in nitesimal increment of time dt is considered, the corresponding in-
nitesimal increment of strain d"fr can be expressed using the di erential form of
(4.3):
" #
@T
d"fr = (Tg )  dr : (4.4)
@r T =Tg

83
Chapter 4 4.1. Frozen-in Strains
It is desirable to minimise generated frozen-in strain increments "fr . They can
be zero only if the temperature gradients at solidi cation are zero. Such case is
called a uniform solidi cation process. This is a highly idealised case, where all
points in the material solidify at the same instant, which is almost impossible in the
practice. However, in any real solidi cation cases there are points which solidify in a
uniform manner. The temperature gradients in points which solidify last locally are
zero, and no strain is frozen-in that region. All other solidi cation processes can be
classi ed as non-uniform solidi cation processes. To conclude, the frozen-in strains
are induced into polymeric materials during non-uniform solidi cation processes and
are inevitable in practice.
The second term in Equation (4.4) is a vector eld with its elements equal to the
temperature gradients at the moment of solidi cation for any material point r. If
both sides of the Equation (4.4) are the total di erentials of appropriate functions,
we obtain: " #
Z Z
@T @T
"fr (r) =  dr = @r  dr; (4.5)
@r T =Tg T =Tg
C (r) C (r)
and nally:
"fr (r) = Tres(r): (4.6)
C (r) is the integration path that need to be de ned for every point in the material r.
As already mentioned in Chapter 2, the frozen-in strains can be described in terms
of a residual temperature eld Tres, Equation (4.6). When this residual temperature
eld is applied as an actual temperature distribution in a thermo-elastic analysis,
the residual stresses are given by the resulting thermal stresses. The problem is now
reduced to solving the thermal stresses. It remains to study closely the integral term
in (4.5). The following substitution can be introduced:
Z
@T
Tres (r) =  dr: (4.7)
@r
T =Tg
C (r)

84
Chapter 4 4.2. Frozen-in Temperature Gradients { Non-Conservative Vector Field
Equation (4.7) de nes the residual temperature eld. Mathematically, it represents
the contour integral of a vector eld; in this case that of the temperature gradients
at the moment of solidi cation. This eld must be integrated along the integration
path C (r) for every computational point.
The solution of the contour integral of type (4.7) is simple if it is path independent,
as is the case for all conservative vector elds. The vector eld v(r) is conservative if
there is a scalar \potential" function (r) such that its gradient equals the original
vector eld v(r) [60]. The contour integral of a conservative vector eld is now:
Zr2 Zr2 Zr2
d (r)
v(r)  dr = dr = d (r) = (r2 ) (r1 ): (4.8)
dr
r1 r1 r1

The contour integral of a conservative vector eld depends only on the values of
the potential function (r) at the start and the end of the integration path, r1 and
r2 . Any arbitrary integration path, connecting r1 and r2 gives the same value for
the integral. Usually, the simplest path parallel to the coordinate axes is chosen to
reach each point.

4.2 Frozen-in Temperature Gradients {


Non-Conservative Vector Field
The frozen-in temperature gradient eld is usually a non-conservative eld and that
fact complicates the integration procedure. Components of this vector eld are de-
termined while solving the energy equation. We will use the Cartesian coordinates
(x; y; z) to describe the position of each point. The energy equation gives the tem-
perature for all points in time, T = T (x; y; z; t). At each time, the temperature is a
spatially distributed scalar eld T = T (x; y; z). The gradient of this scalar eld is by
de nition (4.8) a conservative vector eld. All points that solidify simultaneously at

85
Chapter 4 4.2. Frozen-in Temperature Gradients { Non-Conservative Vector Field
time t form the current solidi cation front, T (x; y; z)jt=const = Tg . This solidi cation
front is advancing in time and at time (t + dt) the new solidi cation front location is
determined from T (x; y; z)j(t+dt)=const = Tg . The newly solidi ed region is bounded
by the current and previous solidi cation fronts. The existing temperature gradients
in the newly solidi ed region must be recorded for subsequent determination of the
residual temperature eld Tres(x; y; z). While solving the energy equation, for points
satisfying T (x; y; z; t)  Tg and T (x; y; z; t + dt) < Tg , temperature gradients are
\copied" into the frozen-in temperature gradient eld (Figure 4.1). This procedure
is repeated until all points have solidi ed.
Additionally, we can store the solidi cation time for each material point. The new
scalar eld of tsol = tsol (x; y; z) has isolines representing solidi cation fronts. The
isochrone, tsol = t, determines the position of the solidi cation front at time t. At
time (t + dt) the solidi cation front has moved to a new position corresponding to

t
ons
t=c

(t+dt)=const

t sol
'fro
zen
-in
' gr
adi
ent
fiel
d

Figure 4.1: Formation of the frozen-in temperature gradient eld.

86
Chapter 4 4.3. De nition of the Integration Path
the isochrone tsol = (t + dt). Although a section of the conservative vector eld has
been copied into the frozen-in temperature gradient eld at each time interval, the
overall conservativeness of the new eld is not preserved: it is not enough to copy the
vector eld components to preserve the conservative feature of an vector eld, one
also has to preserve the local variations of this vector eld. In the solidi cation case
the conservativeness of the frozen-in temperature gradient eld is disturbed in all
directions apart from that of the solidi cation front. Therefore, the contour integral
from (4.7) becomes path-dependent. This requires the appropriate integration paths
to be identi ed for each solidi cation case in order to obtain a unique solution for
Tres .

4.3 De nition of the Integration Path


In one-dimensional cooling, Figure 2.8, the material solidi es along planes, all par-
allel to each other. The frozen-in strains and other dependent variables only change
in the direction of the coordinate axis orthogonal to the solidi cation planes. In the
other two coordinate directions the changes are zero. In one-dimensional cooling,
the frozen-in temperature gradients are all colinear vectors. The paths of their in-
tegration, when calculating Tres, are lines parallel to the z axis. This direction also
coincides with that of principal heat ow lines during solidi cation, which are always
orthogonal to the solidi cation front T = Tg . For this case the contour integral (4.8)
reduces to a simple scalar function integral that can be integrated in a conventional
way:
Zr2 Zz2
v(r)  dr = v (z )dz: (4.9)
r1 z1

The same principle can now be applied for determining integration lines in the
multidimensional case. In the most general cooling case, solidi cation fronts are
87
Chapter 4 4.3. De nition of the Integration Path
no longer planes, but curved surfaces in space. The principal variation of frozen-
in strains is again in the direction of the principal heat ow lines at solidi cation.
This direction is also orthogonal to the solidi cation front. If we assume that the
integration paths are lines always orthogonal to the solidi cation front surface, there
is a total analogy between the one-dimensional and arbitrary cooling cases (Figure
4.2). Physically and intuitively, this is the right choice for integration paths, since
one would expect the frozen-in strains to be generated between successively solidi ed
layers. For neighboring points on the same solidi cation front there is no additional
frozen-in strain since the di erential temperatures between them at the moment of
solidi cation are zero.
The family of lines which are orthogonal to the solidi cation fronts are unique in
every solidi cation case. This ensures the unique solution for Tres(x; y; z). The di-
rection orthogonal to the solidi cation front also coincides with the direction of the
frozen-in temperature gradient vectors. The integration path through any point is
no longer a straight line but a curve in space. The tangent on it in any point along
is always parallel to the frozen-in temperature gradient direction. The orientation of

Figure 4.2: Integration lines orthogonal to solidi cation fronts: a) 1-D cooling case,
b) the multidimensional cooling case.

88
Chapter 4 4.4. Features of the Orthogonal Lines
the integration path dr is chosen to point in the direction of the advancing solidi ca-
tion front. That way each integration path points towards the \hotter" region of the
material, the region that will solidify later. Such orientation is also convenient since
it always produces positive residual temperature increments along the integration
path:

@T
@r T =Tg
 dr > 0: (4.10)
Both vectors in Equation (4.10) are pointing the same way. The integration paths
are referred to as the \orthogonal lines" hereafter.

4.4 Features of the Orthogonal Lines


Orthogonal lines are integration lines for obtaining the residual temperature eld and
they are always orthogonal to the solidi cation front. These are a unique family of
lines for any solidi cation case. Physically, they can be considered as heat ow lines
during solidi cation. There are several important features speci c to orthogonal
lines present in almost every solidi cation case. It is useful to examine these features
more closely.

4.4.1 Orthogonal Lines at Symmetry Planes


There are many cooling cases with symmetry planes. When analysing a at plate
we often have cases with through-thickness symmetry, often the conditions on the
right and left hand side of the plate are identical, etc. In such cases we only need
to determine the dependent variables for a smaller segment of the plate surrounded
by the symmetry planes on several sides. All dependent variables, including the
temperature T , have no variation in the direction normal to the symmetry plane.
This produces a zero temperature gradient in the direction normal to the symmetry
89
Chapter 4 4.4. Features of the Orthogonal Lines
plane. Consequently, the resulting temperature gradient vector lies in the plane of
symmetry, and so does the frozen-in temperature gradient vector. Consequently,
the orthogonal lines will also lie in the plane of symmetry (Figure 4.3). In a two-
dimensional analysis, each line of symmetry represents one orthogonal line.
symmetry
plane
T=Tg
gradT
gradT T=Tg
orthogonal lines
symmetry
gradT plane

Figure 4.3: Orthogonal lines at symmetry planes.

4.4.2 Radially Distributed Orthogonal Lines


Orthogonal lines always follow the heat ow lines at solidi cation. They all start
on cooling boundaries and they point into the warmer interior. As the solidi cation
is progressing, the softened region becomes ever smaller. When solidi cation is
almost complete, the heat is conducted radially from the remaining softened region
[64, 65]. Positive temperature gradients will inevitably point towards the hottest
point in such region. The orthogonal lines will therefore stream towards the point
which locally solidi es last.
Figure 4.4 illustrates one such cooling example. It shows two important steps in the
residual temperature analysis: a) position of the solidi cation fronts in time, i.e.
solidi cation time contours tsol =const, and b) corresponding integration lines. The
material in Figure 4.4 is cooled down from an elevated temperature symmetrically
from all boundary surfaces. As expected, solidi cation starts from the edges and
advances towards the point on the intersection of two symmetry lines A-A and B-B.

90
Chapter 4 4.4. Features of the Orthogonal Lines

A A

B B B

A A
a) Solidification fronts b) Orthogonal lines

Figure 4.4: Cooling example { symmetrical cooling from all sides: a) solidi cation
fronts, b) orthogonal lines.
This is the point that solidi es last and all orthogonal lines stream to it. Its should
also be noticed that the point on the intersection of symmetry lines is usually the
point that solidi es last.

4.4.3 Singular Points in the \Hot" Region


It seldom happens that the cooling method results in near one-dimensional heat
transfer conditions. A typical example is a plate subjected to cooling from its both
major surfaces, i.e. the top and bottom surfaces. In such an example solidi cation
initiates at the boundary surface that is subjected to more intensive cooling. At a
later stage of cooling it will initiate also on the opposite boundary surface. Both
solidi cation fronts are now advancing towards the plate interior. At one instance
they will meet and the whole plate will become solid. Unlike the previous example
where, we had a single point which was last to solidify, here the whole region so-
lidi es simultaneously; this will be termed the \hot" region [64, 66]. As before, the
orthogonal lines are pointing towards the hot region (Figure 4.5).
Figure 4.5 shows the solidi cation fronts and integration lines in one-dimensional
cooling of a plate. The plate is cooled from both the top and bottom surfaces but
91
Chapter 4 4.4. Features of the Orthogonal Lines

weld line

a) Solidification fronts b) Orthogonal lines

Figure 4.5: 1D Cooling example { intensive cooling from the top surface: a) solidi-
cation fronts, b) orthogonal lines.
cooling conditions on the top surfaces are more intensive then on the bottom. This
will result the top half of the plate being colder throughout the cooling and will shift
the hot region downwards. In the region where two solidi cation fronts meet the
frozen-in temperature gradient vectors are pointing the opposite directions. At the
point where the solidi cation fronts meet they are zero. Such points will be called
the singular points, since the direction of the orthogonal line in them is not de ned.
The point which solidi es last locally is another example of a singular point. It is
useful to note that orthogonal lines can only intersect in singular points.

4.4.4 The \Weld" Line


In an arbitrary cooling case we can nd almost all of the orthogonal line features
previously mentioned. One such cooling example is shown in gure 4.6. In this
example the cooling is concentrated in the middle of the top surface. Additionally,
heat losses are allowed on all remaining boundaries. The solidi cation starts from
all boundaries but is most intense in the region close to the top surface. The solidi-
cation completes in points P1 and P2 . In this example we can notice the following
features:
 Orthogonal lines in the vertical symmetry plane A-A,
92
Chapter 4 4.4. Features of the Orthogonal Lines

Figure 4.6: Arbitrary cooling example: a) solidi cation fronts, b) orthogonal lines.
 Singular point on line A-A where the solidi cation fronts from top and bottom
side coalesce,
 Two orthogonal lines intersecting in the singular point,
 Orthogonal lines stream radially towards points that solidify last locally, e.g.
points P1 and P2 .
The orthogonal lines in this example appear to converge to a \hot" or weld line
instead of the hottest points in the domain P1 and P2 . This is merely apparent
since the weld line streams towards the points that solidify last also. The weld line
is a special case of an orthogonal line which exists only in the hot region [64]. At
the moment of solidi cation, the solidi cation front coincides with the isothermal
surface T = Tg . The temperature gradient vectors are normal to the solidi cation
front and determine direction of the integration line. The magnitude of this vector
is proportional to the heat conducted across the solidi cation front: the more heat
is conducted through the solidi cation front, the faster the solidi cation front ad-
vances. On the same solidi cation front the conducted heat is not uniform, but it
varies from point to point. If there is a point which conducts the minimum heat,
the solidi cation front advance will be slowest in this point. This apparent \lag"

93
Chapter 4 4.4. Features of the Orthogonal Lines
behind of the solidi cation front is seen in the orthogonal line family as a weld line
(Figure 4.7).

Figure 4.7: Weld line in the cooling domain.


The weld line connects the points on di erent solidi cation fronts where the frozen-
in temperature gradient vectors have the lowest magnitude. Physically, the weld line
divides the material domain into two regions; during solidi cation the heat cannot
ow to the region on the opposite side of the weld line. The determination of the
weld line will be an important step in the numerical procedure for determining the
orthogonal lines.

94
Chapter 5
Finite Volume Residual Stress
Analysis
Previously, we described the generation mechanism of solidi cation-induced residual
stresses in polymeric materials. The frozen-in strains have been identi ed to be the
main reason for the presence of residual stresses after thermal processing. These
strains have been linked to the thermal history of the material and later described
in terms of the residual temperature eld Tres. The response of the material to
frozen-in strains is identical to the response to the residual temperature eld. The
resulting residual stress distribution has been found to be mathematically complex
even for one-dimensional cooling cases and hence numerical methods are preferred.
This Chapter brie y explains one such numerical method for determining residual
stresses in the material.

5.1 Governing and Constitutive Equations


The discipline of solid mechanics describes the deformation behavior of solid bodies
subjected to systems of forces. The forces acting on the body can be generated in
many di erent ways: mechanically, thermally, result from inertia, etc. The material
responds by deforming to any combination of these forces. There are fundamental

95
Chapter 5 5.1. Governing and Constitutive Equations
laws of physics which any substance in nature, including solid bodies, obey. This
laws will be applied here to a material on the macroscopic, continuum level, and
thus resulting the governing or conservation principles. The concept of continuum
assumes the length-scales to be substantially larger then the characteristic length
scales of the discrete structure, i.e. the average molecular free path. Under this
assumptions, the material lls the space continuously and all physical properties
become mathematically continuous functions in space, i.e. elds [58].
In mechanics of solid bodies the most important governing equations are the con-
servation of linear momentum and the conservation of energy, also known as the
2nd Newton's law of motion and the 1st law of thermodynamics. The latter has al-
ready been mentioned and discussed in Chapter 3 as the energy equation. Newton's
second law, i.e. the momentum equation, can be expressed in the following integral
form [57]:
Z Z Z
d
v dV = n   dS +  fb dV ; (5.1)
dt
V S V
inertial forces surface forces body forces

where V is the volume of the arbitrary part of the body, S is the corresponding
bounding surface with the outward pointing unit normal vector n, t is the time, v
is the velocity vector,  is the symmetrical stress tensor,  is the density and fb
is the total body force. The left hand side of Equation (5.1) represents the rate
of change of linear momentum which is equal to the sum of all surface and body
forces acting on and in the body. The operator \ d" in Equation (5.1) represents the
material derivative. For solid materials undergoing small dimensional changes the
momentum equation is simpli ed into:
Z Z Z
@ @u
 dV = n   dS +  fb dV: (5.2)
@t @t
V S V
The newly introduced vector eld variable u represents the displacement vector
of the solid material. Equation (5.2) does not uniquely de ne the unknown dis-
96
Chapter 5 5.1. Governing and Constitutive Equations
placement u and stress eld . Mathematically, it represents a system of 3 inte-
gral/di erential equations with 9 unknown eld components, i.e. 3 unknown dis-
placement vector components and 6 unknown stress tensor components. Additional
equations are therefore required to close the system. These subsidiary equations
need to relate the unknown quantities from the governing equations and are better
known as the constitutive equations. They are phenomenological and describe the
behavior of a speci c material. For a solid material the constitutive equation usu-
ally relates the stress tensor components to appropriate measures of strain. Hooke's
law represents one such constitutive equation applicable for linear isotropic thermo-
elastic materials. Hooke's law can be expressed as follows [57]:
 
 = 2" + tr " 3K (T T0 ) I: (5.3)
Here, " is the strain tensor, T and T0 are the real and reference temperatures
respectively, where T0 corresponds to thermally undeformed state; K is the bulk
modulus of the material, is the coeÆcient of linear thermal expansion,  and 
are Lame elastic constants, \tr" is the trace 1 operator and I is the identity tensor.
The Lame elastic constants and the bulk modulus are related to the more familiar
elastic constants, E and  , the Young's modulus and the Poisson's ratio [67]:
E E E
= ; = ; 3K = 2+3 = (5.4)
2(1 +  ) (1 +  )(1 2 ) 1 2 :
Hooke's law in the form (5.3) is applicable to most general linear elastic stress
analysis cases, including the idealised one-dimensional cases involving in nite plate,
mentioned previously in Chapter 2.
As an illustration of the use of the general Hooke's law, we will evaluate it for the
in nite plate geometry shown in Figure 2.1. The speci cs of this case are as follows:
"xx = "yy = "; xx = yy =  and zz = 0: (5.5)
2 3
"xx "xy "xz
1 tr " = tr 4"yx "yy "yz 5 = "xx + "yy + "zz
"zx "zy "zz

97
Chapter 5 5.1. Governing and Constitutive Equations
When the conditions (5.5) are substituted into Equation (5.3), for the normal stress
components we obtain:
2 3 2 3 2 3 2 3
 " " + " + "zz (T T0 )
4 5 = 2 4 " 5 +  4" + " + "zz 5 3K 4 (T T0 )5 : (5.6)
0 "zz " + " + "zz (T T0 )
After substituting "zz from the third row equation in (5.6) into any of the remaining
two equations, the in nite plate constitutive equation (2.3) is recovered:
=
E  
" (T T0 ) ; with "zz =
2 "+ 1 +  (T T ): (5.7)
1  1  1  0

In another one-dimensional case relevant to the in nite plate case, we can assume
that the plate is fully constrained in one of the two plate directions, e.g. in the
y -direction (Figure 2.1). This would result in plane deformations in the y -direction
for which "yy = 0. We will refer to this case as one of \plane strain" cases. With
xx =  and "xx = ", Hooke's law now becomes:
2 3 2 3 2 3 2 3
 " " + 0 + "zz (T T0 )
4yy 5 = 2 4 0 5 +  4" + 0 + "zz 5 3K 4 (T T0 )5 : (5.8)
0 "zz " + 0 + "zz (T T0 )
The plane strain in nite plate constitutive equations are:
 
=
E "
(T T0 ) ; with "zz =

"+
1 +  (T T ):
1  1+ 1  1  0

(5.9)
Both versions of the Hooke's law derived for in nite plate geometries can be used
for validation of numerical stress solutions.
The tensorial equation (5.3) brings 6 new equations to our system of equations,
but also introduces 6 unknown strain tensor components of ". The system of the
governing momentum equation can be closed by de ning the strain tensor " purely
in terms of the displacement vector components u. If the strains in the system are
small, the strain tensor is de ned as:
1
" = ru + (ru)T :

(5.10)
2
98
Chapter 5 5.1. Governing and Constitutive Equations
Nabla (r) is the gradient operator and ru is the gradient of a vector u resulting
in a second rank tensor [58]. The term (ru)T is the transpose tensor of ru. By
substituting the strain tensor into the Hooke's law equation (5.3) we obtain:
  
 =  ru + ruT + tr(ru) 3K (T T0 ) I: (5.11)
Since in the residual stress analysis the changes of displacements in time are very
small, all inertial e ects can be neglected in (5.2) by setting the time derivative term
to zero. This will produce the steady state equation for u:
Z Z
n   dS +  fb dV = 0;
S V
and by using Equation (5.11) we obtain:
Z Z
   
n  ru+ru + trru 3K T I dS +
T
 fb dV = 0: (5.12)
S V
Here, the reference temperature T0 is assumed to be zero and will be used here-
after. If we assume that temperatures in the system T are known, Equation (5.12)
represents a closed system of vector equations, with the single unknown variable
u. To obtain the unique steady state solution for u in any material point r, the
conditions along the domain boundary rB need to be speci ed. The most commonly
used boundary conditions for the momentum equation are [68]:
 Fixed displacement boundary. Displacement vector u is speci ed at the do-

main boundaries rB, i.e. u r=r = uB.
B

 Traction boundary. Components of the traction vector (n  ) are prescribed


at the domain boundaries. A free surface is a special case of this boundary
condition where the traction is equal to zero.
 Mixed boundary. Prescribed is the normal component of the boundary dis-
placement and the tangential component of the traction vector:
(uB)n = un and (n  )t = t :
99
Chapter 5 5.2. Fundamentals of the Finite Volume Method
The symmetry plane condition ts into this category with the zero normal
displacement component, un = 0, and zero shear traction, t = 0.
As in most cases, the steady state momentum equation (5.12) is a mathematically
very complex equation. It is the di erential/integral equation for which the analyt-
ical solution often cannot be expressed in a closed form. The momentum equation
can be expressed in purely di erential form, but rst the surface integral term must
transform into the volume integral by using the Gauss theorem [57]:
Z Z
   
r  ru+ru + trru 3K T I dV +  fb dV = 0:
T
(5.13)
V V
Since the volume V in (5.13) is arbitrary, the sum of integrals on the left hand side
can be equal to zero only if the sum of all integrands is always zero. This yields the
di erential form of the momentum equation:
   
r  ru + ruT + trru 3K T I +  fb = 0: (5.14)
This is the momentum equation expressed in the classical form. Mathematically it
represents the system of 3 partial, second order di erential equations. The solution
for displacement vectors u is usually obtained by appropriate numerical methods.
In the following Section the Finite Volume method (FVM) is introduced. It is a
numerical method capable of solving our momentum and energy equations.

5.2 Fundamentals of the Finite Volume Method


Physical problems in continuum mechanics are described with the mathematical
expressions, which are combined together to form the mathematical model. For
multidimensional physical phenomena a mathematical model is usually a system of
integral or di erential equations which are complex to solve. For a well posed phys-
ical systems a unique solution to the problem exists. The only remaining problem
is to nd this solution.
100
Chapter 5 5.2. Fundamentals of the Finite Volume Method
When classical mathematics cannot provide us with the solution, approximate nu-
merical methods are employed. The aim of the numerical methods is to trans-
form the equations of the mathematical model into simpler, solvable mathematical
schemes. The simpli ed objects are usually systems of algebraic equations. One
such method with the strong physical basis is the Finite Volume method. The -
nite volume method is suitable for solving the integral governing equations such as
(5.13), without having to go deeper down to the di erential level (5.14), preferred
by the mathematicians. The integration volumes V in the governing equations (3.1)
and (5.1) can be any arbitrary part of the physical domain. The main objective of
the nite volume method is to conserve the basic physically conserved quantities in
a number of prede ned volumes forming the physical domain.
Sf
f
N

P
z
r
x y
CV

Figure 5.1: A control volume of an arbitrary shape.

In the nite volume method the domain is subdivided into a nite number of control
volumes (CV) or cells of volume V , bounded by a number of at surfaces Sf or cell
faces (Figure 5.1) [68, 69, 70, 71, 72]. This process is known as a space discretisa-
tion. Each cell is surrounded by a number of neighbouring control volumes. Two
neighbouring cells P and N exchange information through a common cell face f .
The computational point P is placed in the centroid of the control volume or cell

101
Chapter 5 5.2. Fundamentals of the Finite Volume Method
centre, and on the boundary cell face the equivalent point is located in the centre
of the face. The coordinates rP and rf satisfy the centroid conditions:
Z Z
(r rP)dV =0 and (r rf) dS = 0:
VP Sf

In each cell, any unknown physical quantity  is assumed to have a linear variation
in space around the cell centre P . At any point r within the cell,  is calculated
[73]:
(r) = P +(r rP)  (r)P : (5.15)
With the linear variation, the average quantity of  in each control volume is that
in the cell centre P . Consequently, the average of  on each cell face Sf equals
that in the face centre f. In the centre of the shared face between two control
volumes, the average physical quantity f must be equal regardless from which side
of the interface it is calculated, P or N. This simple requirement is one of the main
features of the nite volume method, and it ensures the conservation of the physical
quantities throughout. Physical properties must not be dissipated on the cell faces,
the averaged properties must be \transported" from one side of the interface to the
other side in its entirety.
Like the space domain, the time domain is also discretised. The total time interval
t is subdivided into arbitrary number of time steps Æt. The distribution of any
physical property is again assumed to be linear within the time step:
 
@ m
(t + Æt) = (t)+ @t
Æt: (5.16)
The superscript m indicates the current time interval t. With the time and space dis-
cretisation in place, governing equations will be discretised for each control volume
and in each time step.

102
Chapter 5 5.3. Discretised Forms of the Governing Equations
5.3 Discretised Forms of the Governing Equations
This Section outlines the conversion of the mathematically complex governing equa-
tions into a much simpler systems of algebraic equations. Each individual term in
the governing equations will have the simpli ed nite volume substitute expression.
This conversion process is known as the equation discretisation [74].
For the proposed residual stress analysis, the temperatures will be obtained by
solving the energy equation in time and space domains, T = T (r; t), while the
momentum equation needs solving only in the space domain u=u(r). This two
equations are given again (note the residual temperature Tres in place of the actual
temperature T in the momentum equation):
Z Z
@
@t
cT dV = n  (krT ) dS; (5.17)
VP S
Z
   
n  ru + ruT + trru 3K Tres I dS = 0: (5.18)
S

In the present work, body forces and heat sources are omitted since they are negligi-
ble compared to the remaining terms. In the nite volume discretisation, the volume
integrals are approximated with the simpler volume integrals, and the surface in-
tegrals are approximated with the simpler surface integrals. Due to the assumed
linear distribution of the integrands in the control volume, the mid-point rule can
be employed, and we obtain for the integrals:
Z
cT dV = (cT )mP VP; (5.19)
VP
Z n
X
n  (krT ) dS = nf  (krT )mf Sf ; (5.20)
S f=1

103
Chapter 5 5.3. Discretised Forms of the Governing Equations
Z
   
n  ru + ruT + trru 3K Tres I dS =
S
n
X    
= nf   ru + ruT + trru 3K Tres I f Sf: (5.21)
f=1

Likewise, for the time derivative term we obtain:


Z
@
@t
cT dV = VÆtP (cT )mP (cT )mP 1
: (5.22)
VP

The fully implicit time discretisation of the energy equation is performed in Equa-
tions (5.19) and (5.20). Such a discretisation technique uses the integrand values
from the current time step m to evaluate the integrals [67]. The n in previous equa-
tions denotes the number of nearest neighbours for the cell P, neighbours that share
faces 1,2,...,n with cell P. All quantities that need to be calculated on the cell faces
can be interpolated from the values in cell centres that share the face. If the mate-
rial properties are constant they do not need to be interpolated. However, material
properties are usually temperature-dependent or vary in space. If the variation of
these properties is not signi cant between two neighbouring cells, linear interpola-
tion can be employed. In case of large variations, i.e. two di erent materials, the
harmonic interpolation would produce more appropriate material properties [75, 76].
The remaining unknowns in the discretised equations are functionals of the unknown
variables T and u, interpolated in the face centres. They are (rT )f; (ru)f; (ruT)f
and (tr ru)f 2. The idea is to approximate these terms using the cell centre value
of P and the cell centre values in the nearest neighbours only. In the face inter-
polation of each functional, the contributions from the cell centres involved need
to be consistent with the employed spatial discretisation, primarily the geometrical
parameters such as distances from cell centres to the face centres, orthogonality of
2 (tr ru) = (r  u)

104
Chapter 5 5.3. Discretised Forms of the Governing Equations
control volumes [73], etc. Two typical geometrical arrangements around the cell
faces are shown in Figure 5.2.
nf
k
o
dPN nf dPN nf

P f N P f N

(a) orthogonal mesh (b) non-orthogonal mesh

Figure 5.2: Orthogonal and non-orthogonal cell face topologies.


As an example, gradient terms will be discretised for both orthogonal and non-
orthogonal face topologies. In both cases we only need the projection of the gradient
in the normal direction nf for Equations (5.20) and (5.21). From the neighbouring
cell centres we can estimate the gradient component in the direction of the distance
vector dPN which connects the cell centres P and N { the orthogonal part of the
gradient. In the orthogonal case (Figure 5.2a) the vector dPN and the face normal
nf are colinear. The component of the gradient required for discretisation is the
orthogonal gradient only:
TN TP uN uP
nf  (rT )f = and nf  (ru)f = (5.23)
j dPNj j dPNj :
In the non-orthogonal case (Figure 5.2b) the distance vector is no longer colinear
with the normal face vector. The required projected gradient can be decomposed
into two components, i.e. the proportion of the orthogonal gradient that can be
accurately estimated from P and N and the cross-gradient component:
TN TP
nf  (rT )f = (o + k)  (rT )f = j oj + k  (rT )f; (5.24)
j dPNj
uN uP
nf  (ru)f = (o + k)  (ru)f = j oj + k  (ru)f: (5.25)
j dPNj
105
Chapter 5 5.3. Discretised Forms of the Governing Equations
The choice of the orthogonal vector o (colinear with the distance vector dPN) and
the correction vector k is not uniquely de ned; however their sum must equal the
unit face normal face vector nf, Figure 5.2b. Vectors o and k mainly in uence
the numerical convergence. Several approaches of their selection are discussed by
Jasak [73]. The temperature and displacement gradients on the cell face f , from
the right hand sides of Equations (5.24) and (5.25), are calculated explicitly from
latest values of temparatures and displacement vectors [67, 73]. The non-orthogonal
discretisation turns into orthogonal on orthogonal computational meshes since we
have j oj = j nfj = 1 and k = 0. When discretising the at plate geometry, which is
the case in present work, it is natural to create a fully orthogonal spatial domain for
which the non-orthogonal correction is always zero. The remaining functionals of T
and u are approximated in a similar manner and will not be discussed here. If during
the discretisation we come across cell faces located at domain boundaries, the un-
known face functionals are approximated from the given boundary conditions. The
discretisation procedure of boundary faces is identical to discretisation of internal
faces.
It is important to note that expressions (5.24) and (5.25) are algebraic equations
and so are all other nite volume substitutes for di erent terms from the momentum
and the energy governing equations. When we combine all discretised terms for each
control volume we obtain the algebraic equations in the following form [67]:
n
X n
X
aP TP aK TK = b and P uP K uK = : (5.26)
K=1 K=1

From the discretised Equations (5.26) it is clear that the cell temperature and the
displacement vector are directly in uenced by the local temperature and the dis-
placement elds. If they are known in n nearest neighbours, the values in the cell
centre TP and uP are uniquely determined from it. But, Equation (5.26) represents
a set of coupled equations and neither the N cell centre temperatures nor the dis-
106
Chapter 5 5.3. Discretised Forms of the Governing Equations
placement vectors are known in advance. All coeÆcients in (5.26) are determined
during the discretisation. The coeÆcients are constant if the material properties are
not dependent on the solution itself, i.e. temperature independent thermal proper-
ties in the energy equation. Generally, this is not the case and original linear system
of equations (5.26) becomes an non-linear system that must be solved iteratively.
For every control volume, equations of the form (5.26) can be assembled thus creating
a systems of equations, i.e. one scalar system for temperatures and one vector system
of equations for displacements:
[a] [T ] = [b] and [ ] [u] = [ ]: (5.27)
Here [a] and [ ] are sparse matrices containing discretisation coeÆcients, [T ] and [u]
are single column matrices containing unknown dependent variables for all control
volumes, and [b] and [ ] are single column matrices containing all source terms. If the
total number of cells is N , we have to resolve N unknown temperatures within each
time step and 3N unknown steady state displacement components for 3-D problems
(2N for 2-D problems). The scalar system can be solved with various available
linear equation solvers. However, the vector equation needs some attention before
the solver. It can be solved with several di erent procedures. The most commonly
used solution procedures are [77]:
 Simultaneous solution procedure.
All displacement vector components are calculated simultaneously in the same
system of equations. The vector system is transformed into a scalar system
with 3N unknowns. When fully populated, the coeÆcient matrix [a] of that
system becomes of the order (3N  3N ) matrix. For very large systems the
size of this matrix becomes substantial making it impractical for handling
on computers. Additionally, the coeÆcient matrix is no longer diagonally
dominant, which make the system of linear equations more diÆcult to solve.
107
Chapter 5 5.4. The Iterative Segregated Solution Algorithm
 Segregated solution procedure.
The vector system can be transformed into 3 subsystems of equations by \pro-
jecting" all vector equations in the coordinate axis directions. This would yield
a scalar system of equations for each direction, i.e. the displacement vector
components ux; uy and uz . There will be 3 coeÆcient matrices [a], each of
the order (N  N ). Each individual system now contains the remaining two
vector components in it, producing a non-linear system of equations. The
inter-equation coupling here is achieved via the source terms by assuming the
remaining displacement components to be known. In the segregated procedure
the equations are solved iteratively. Due to a smaller matrices, this method
substantially reduces the required computing resources. Also, the obtained
coeÆcient matrices possess favourable diagonal dominance.
The segregated procedure is favourable in nite volume calculations since it is mem-
ory eÆcient method and handles non-linearities with ease [68].

5.4 The Iterative Segregated Solution Algorithm


The derived systems of equations are almost always non-linear. Non-linearity arises
mainly from the inter-equation coupling and also from the non-linear material prop-
erties. Even with the non-linear systems of equations, we can still use the available
linear equation solvers to obtain the solution. But rst, the non-linear systems of
equations needs to be linearised by temporarily assuming that all remaining depen-
dent variables, apart from the one that we are currently solving for, are known and
constant. A good guess for the remaining variables is usually taken from the initial
conditions or previous iteration results. By using these values we can assemble the
coeÆcient matrix [a] and the source vector 3 [b] and treat them as constants. This
3 Term \vector" in matrix calculus = a single column matrix.

108
Chapter 5 5.4. The Iterative Segregated Solution Algorithm

Figure 5.3: Flow diagram for solving non-linear energy equation system.
results in a system of linear algebraic equations that can be solved in conventional
ways. This process is repeated in succession for all dependent variables until a
converged solution is obtained.
Figure 5.3 shows the sequence of solution steps when solving the energy equation.
The energy equation system becomes non-linear due to temperature-dependent ther-
mal properties. The consequence is that the coeÆcient matrix and the source vector
are not constant, but a function of the current temperature T , which is not known,
[a] = [a(T )] and [b] = [b(T )]. However, thermal properties can be obtained for ap-
proximate, guessed temperatures. After solving this system, a better estimate for
the temperature is obtained and used to update the thermal properties.
109
Chapter 5 5.4. The Iterative Segregated Solution Algorithm

Figure 5.4: Flow diagram for solving decoupled momentum equation system.

110
Chapter 5 5.4. The Iterative Segregated Solution Algorithm
The ow diagram for the solution of the momentum equation is shown in Figure
5.4. Firstly, the ux system is assembled by assuming uy and uz to be known. After
solving this system the new estimate for ux is obtained. The process is repeated for
the uy and uz systems of equations. The convergence of this system depends on the
size and numerical complexity of the problem.
Sometimes simple physical problems lead to very complex numerical systems. A
typical example is a beam geometry subjected to bending loads. The numerical
complexity of the beam case arises from two sources:
 The rate in which the
High geometrical aspect ratio, i.e. length/thickness ratio.
boundary information is exchanged between two opposite domain boundaries
is reduced due to the long distance that boundary information has to travel
(huge number of CVs between the boundaries), and the cascade nature of the
iterative solvers [74].
 High source term non-linearity. The principal variation of the displacement
component (displacement gradient) is not in the direction of its coordinate
axis. The source term in one direction is very sensitive to the assumed dis-
placement components in the other two directions. For example, in the simple
cantilever beam bending case (Figure 5.5) the displacement uy predominantly
changes in the x-direction. The term @uy =@x will be treated as a source term
in the system of equations for ux thus creating a dominant, solution dependent
term in the equation [74].
It is common in the nite volume method to use special numerical procedures to
enhance the numerical convergence of the momentum equation. These are known as
the acceleration procedures. The most widely employed is the multigrid acceleration
procedure [74]. The use of multigrid acceleration greatly reduces the beam bending
diÆculties.
111
Chapter 5 5.4. The Iterative Segregated Solution Algorithm
y
∂u y ∂u y
x >>
∂x ∂y F

Figure 5.5: Cantilever beam in bending { problem with the convergence.


It is typical of the iteration solution procedure that the convergence slows down
signi cantly after a few initial iterations. The solution error is a function that
is zero on the domain boundaries and is increased within the domain, with its
shape dependent on the assumed initial displacement eld [67]. Initial displacements
are usually zero. In the stable convergent solutions these errors are the highest
initially. Since the coeÆcient matrix is sparse, the solution is improved within
each iteration only from the neighbouring cell centres. Therefore, the error pro le
gradually becomes a smooth pro le. The error reduction rate is heavily decreased
when the error pro le becomes smooth since it is calculated locally. The same
error pro le would appear less smooth on a coarser computational grid. If the same
case is modeled on coarser computational grid the error reduction rate for it would
therefore increase. It is also bene cial in the coarse grid case that the number of
control volumes has decreased and the numerical information can propagate faster
throughout the domain. The multigrid acceleration technique combines ne and
coarse grid systems of equations to obtain the converged solution. It utilises all
the bene cial features from the coarse and the ne grid systems that accelerate the
solution convergence, i.e. substantial reduction of errors on the coarse grid and
smoothing the solution on the ne grid.
By using the multigrid acceleration the segregated solution procedure for solving
the momentum equation becomes a very eÆcient solution algorithm. The remain-

112
Chapter 5 5.5. The Orthogonal Line Search Procedure
der of this Chapter describes the numerical algorithm for determining the residual
temperature eld Tres.

5.5 The Orthogonal Line Search Procedure


The residual temperature values Tres required for the momentum equation (5.18)
must be determined for all control volume centre points P. This process requires
the determination of orthogonal lines for each cell centre followed by the numerical
integration of the frozen-in temperature gradients along the orthogonal lines. The
whole process can commence as soon the solidi cation is complete.
The numerical algorithm for determining the orthogonal lines will now be outlined.
It accommodates all speci c features of the nite volume spatial discretisation as
the nite volume method is chosen to carry out the thermo-mechanical analysis.
The numerical orthogonal line is chosen to be a continuous line consisting of a
number of straight-line segments. An orthogonal line enters each control volume in
a point on one of its bounding cell face planes and exits in the point on another face.
The required line segment within each cell simply connects these two points. The
orientation of the line segments is a function of the frozen-in temperature gradients,
recorded for all cell centres while solving the transient energy equation [32].

5.5.1 Cell Face Intersection Points


Let us assume a straight line in 3-D space through a point r0 in the frozen-in
temperature gradient direction b (Figure 5.6). All the points on the line r satisfy
the following parametric vector expression [78]:
r = r0 +  b: (5.28)

113
Chapter 5 5.5. The Orthogonal Line Search Procedure

Figure 5.6: The straight line in 3-D intersecting with the face plane.
Here  is a scalar parameter.  is equal to zero in the reference point r0, it is
positive in the direction of the vector b and negative in the opposite direction. The
straight line (5.28) now intersects with the cell face plane. All in-plane face vectors
are normal to the face normal vector nf giving:
(r rf )  nf = 0; (5.29)
with rf being the position vector any point in the cell face, i.e. face centre. By
equating the last two equations, the intersection point is located via the parameter
 as follows:
(r r )  n
 = f 0 f: (5.30)
b  nf
If we assume a straight line going through the cell centre P in the direction b, then
this line will cross all cell face planes (except planes parallel to the line). All these
intersection points can be located using Equation (5.30). Of all intersection points
we are only interested in the point where the line exits the current cell. This point
is the new point on the orthogonal line that we are following. At this intersection
point we place a new straight line in the new frozen-in gradient direction. The
114
Chapter 5 5.5. The Orthogonal Line Search Procedure
algorithm is repeated until a point on the boundary or point within the solution
domain that solidi es last is reached. For clarity, the algorithm is illustrated on a
two-dimensional FV computational molecule.

Figure 5.7: Orthogonal line tracking procedure.

In Figure 5.7 a straight line is placed at point \2". The line crosses the cell face
planes S1 , S2 , S3 and S4 at points 1, 2, 3 and 4 respectively. It exits the cell P at
point 4 on the cell face S4 . Here the orthogonal line enters the south neighbouring
cell S. Point 4 is the new point on the orthogonal line. Mathematically, this point
can be identi ed from the calculated parameters 1 ; 2; 3 and 4. Among them, 4
is the smallest positive real number [79].

115
Chapter 5 5.5. The Orthogonal Line Search Procedure
5.5.2 The Direction Vector b for the Cell
The initial direction of the orthogonal line from any cell centre P coincides with the
frozen-in temperature gradient direction b. When the orthogonal line reaches the
rst cell face plane, the new direction vector b for the line must be calculated. The
new direction vector b is obtained by interpolating the existing direction vectors
at surrounding cells centres. The interpolation procedure for determining the new
direction vector b consists of two steps:
 Direction vectors are interpolated into corner points of each control volume
(vertex points) from the values at neighbouring cell centres that share the
same vertex point. This can be implemented using the linear Least Square
method [80].
 Linear interpolation on the cell face from the values in vertex points.
The main bene t of the interpolation procedure described above is that it is general
and can handle arbitrary mesh topologies.
In some solidi cation cases the proposed algorithm can su er from numerical errors.
If the orthogonal lines are very long they cross many control volumes from the
starting to the nishing point. The numerical errors accumulate along the line and
may become signi cant. The direction vector b used to cross the cell P can be
estimated more accurately. If the direction vector b for the current cell is taken as
a direction in the midpoint of the current line segment, it would produce a second-
order accurate solution. The midpoint direction is interpolated from the directions
at the beginning and end points of the segment. Unfortunately, the middle point
of the segment is not known in advance. Several predictor/corrector cycles can be
used to locate the end point of each segment, and consequently its midpoint as well
(Figure 5.8).
116
Chapter 5 5.5. The Orthogonal Line Search Procedure

Figure 5.8: Second-order accurate line search procedure.


Figure 5.8 illustrates one such iteration sequence to obtain the second-order accurate
orthogonal line segment. The direction of the line segment is gradually improved
from b1 to b2, then to b3 ... until the end point of the segment stops changing.
The direction vectors in the starting point bS and in all calculated end points bE1 ,
bE2 ... are interpolated from vertex points. This procedure typically converges
in 2-3 iterations to a very low tolerance. However, due to the nite resolution of
the mesh, some numerical instabilities can occur in solidi cation cases. These are
addressed in the following Section.

5.5.3 Numerical Determination of the Weld Line


During the computation of orthogonal lines the numerical diÆculties can arise in re-
gions where the frozen-in temperature gradients signi cantly change their directions
[64]. The numerical algorithm is unstable, thus producing orthogonal lines which
are physically not meaningful. Therefore, the orthogonal lines should not be allowed
to enter such regions during the search procedure. A several typical situations are
presented in Figure 5.9. In a smooth direction eld, Figure 5.9a, the method is
very stable and convergent. If the direction vectors point in the opposing direc-

117
Chapter 5 5.5. The Orthogonal Line Search Procedure

Figure 5.9: Orthogonal lines determined for various direction elds:


a) smooth direction eld, b) alternating direction eld, c) opposing direction eld.
tions across the cell then unacceptable \zig-zag" patterns are obtained, Figure 5.9b.
In the extreme case of one-dimensional solidi cation, the direction vectors across
the cell are all colinear but point in opposite directions. The orthogonal line now
switches between two cell boundary faces inde nitely, Figure 5.9c. These potentially
dangerous regions were identi ed in the Chapter 4 as hot regions. Physically, the
event shown in Figure 5.9b occurs when there is a point on the solidi cation front
118
Chapter 5 5.5. The Orthogonal Line Search Procedure
with the minimum solidi cation rate, and the event from Figure 5.9c when two so-
lidi cation fronts coalesce. All cells must be inspected for such points prior to the
determination of the orthogonal lines. When detected, such points can be connected
to form the weld line (see Section 4.4.4).

Figure 5.10: Determining the weld line.

If we again assume a two-dimensional FV computational molecule (Figure 5.10), the


problem can be formulated mathematically in the following way. We are searching
all cell faces of a current control volume for points with the mathematical minimum
of the frozen-in temperature gradient. The gradient vectors can be determined along
any cell face by linear interpolation of the frozen-in gradients from the vertex points.
For cell face lying between the vertices V2 and V3 (Figure 5.10) we have:
b = bV2 + (bV3 bV2 ): (5.31)
The scalar is in the range 0   1 on the cell face. The magnitude of vector b

119
Chapter 5 5.5. The Orthogonal Line Search Procedure
is determined from:
j bj2 = bb = (bV2 bV2)+2 bV2 (bV3 bV2)+ 2 (bV3 bV2 )(bV3 bV2 ):
(5.32)
The squared magnitude function j bj 2 has a minimum at the same points as the
magnitude function itself (j bj). By di erentiating (5.32) with respect to we get
points with the minimum:
bV2  (bV3 bV2 )
= (5.33)
(bV3 bV2 )  (bV3 bV2 ) :
If this minimum falls within the face (0   1), this point is marked and will be
used to construct the weld line. By minimising the squared magnitude function j bj 2
Equation (5.33) will additionally identify points with zero frozen-in temperature
gradients and points where two solidi cation fronts coalesce. Therefore the single
expression identi es the weld line criteria, which greatly simpli es the implementa-
tion of the algorithm. With the weld line determined, every time an orthogonal line
attempts to cross the weld line, it is terminated, Figure 5.11. The weld line brings
numerical stability to the whole orthogonal line search procedure and will also help

Figure 5.11: Orthogonal line terminating on the weld line.


120
Chapter 5 5.5. The Orthogonal Line Search Procedure
in determining the integration constant for the orthogonal line.
The numerical description of FV meshes is similar to meshes used by other numerical
methods, such as FE. With only minor adaptations, the presented algorithm can
be used in conjuncrion with other numerical methods. The orthogonal line search
procedure can also be employed to solve other physical and mathematical problems,
calculating streamlines in uids, solving partial di erential equations, etc.

5.5.4 Numerical Integration along the Orthogonal Lines


After the orthogonal lines are numerically determined it only remains to integrate
the frozen-in temperature gradient vectors along them. Since the segments of the
integration paths are always colinear with the frozen-in vector eld, the problem
is reduced to a one-dimensional scalar function integration. If we assume that the
magnitude of the frozen-in gradient vector varies linearly over the short path incre-
ment within each cell, we can use the midpoint rule to carry out the integration.
This is equivalent to employing the trapezoidal rule for numerical integration [80].
The integration constant for the orthogonal line is chosen to give a zero residual
temperature value at the point which solidi es last, E (Figure 5.12a). In the solidi -
cation cases with the weld line as in Figure 5.12b, the orthogonal line is terminated
at point H on the weld line. It is then proceeded to point E via the weld line. Since
many orthogonal lines will converge to the weld line, it is worth while to numerically
integrate the weld line also. This determines the residual temperatures Tres for all
points on the weld line. The integration constant for any orthogonal line is now the
value for Tres in the point of intersection with the weld line H. This completes the
numerical determination of the residual temperature eld Tres.

121
Chapter 5 5.5. The Orthogonal Line Search Procedure

Figure 5.12: Numerical integration along the orthogonal line:


a) orthogonal line terminates at a point that solidi es last, b) solidi cation case with
the weld line.
5.5.5 Optimisation of the Numerical Procedure
The orthogonal line search procedure can take a substantial amount of computa-
tional time on larger meshes. For each cell centre, an orthogonal line needs to be
followed throughout the domain which can be time-consuming. A few simple im-
provements can be introduced to the original algorithm to considerably increase the
eÆciency of the whole procedure.

122
Chapter 5 5.5. The Orthogonal Line Search Procedure
If the orthogonal line is longer it is more likely that it will pass in the vicinity of
other cell centres for which the residual temperature value is also required. A small
tolerance Æ around any cell centre can be set to detect when an orthogonal line
passes close enough to other cell centres. The more cell centres a single orthogonal
line can serve, the less orthogonal lines need to be determined (Figure 5.13). Usually
the longest orthogonal lines start at the cooling boundaries. These cell centres have
the lowest solidi cation time tsol . All cells can be sorted in the ascending order
regarding the solidi cation time and this can be the order in which the orthogonal
lines are determined. In Figure 5.13 only 3 orthogonal lines C1, C2 and C3 cover
almost 50% of all cell centres in the domain.
The number of numerical integration cycles can also be reduced. The integration
constant for each orthogonal line is chosen to produce a zero value integral at the
point which solidi es last, e.g. point E in Figures 5.12a and b. Di erent starting
points have di erent integration constants to be added on when the last point E
or point on the weld line H is reached, Figure 5.12b. Instead, we can start the

Figure 5.13: Optimised orthogonal line search procedure.

123
Chapter 5 5.5. The Orthogonal Line Search Procedure
numerical integration from points E or H. At these points the integration constants
are known. As we are marching towards the starting point on the orthogonal line
and performing the numerical integration, we can copy the current integral values
for all the cell centres that we come across. The residual temperatures for all cell
centres on the current orthogonal line are determined in one numerical integration
cycle. The above recommendations can reduce the orthogonal line computational
time usually by the factor of at least 10.

124
Chapter 6
Validation of the Numerical
Algorithm

The available one-dimensional analytical solutions for residual stress problems in


a at plate are only valid for idealised solidi cation cases. For other, more realis-
tic cases, the analytical solution does not exist. However, although the available
analytical solutions can give useful qualitative information about the problem, full
quantitative information can be obtained using numerical methods. These must also
be applicable for simpli ed solidi cation cases as for realistic ones. To develop con-
dence in using the numerical methods, the analytical solutions must be recovered
when idealised situations are modeled. In this Chapter we test important steps of
the proposed Finite Volume residual stress analysis algorithm. Later, four typical
solidi cation situations are modeled and the results are discussed in detail.

6.1 Validation of the Orthogonal Lines Procedure


The solidi cation-induced frozen-in strains can be determined numerically as soon
as the solidi cation is complete, i.e. the number of solidi ed cell centres equals
the total number of cells in the computational mesh. During calculation of the
frozen-in strains, the determination of orthogonal lines is the most delicate step to
125
Chapter 6 6.1. Validation of the Orthogonal Lines Procedure
perform. Orthogonal lines are unique when all the frozen-in temperature gradients
b are known. The use of this vector eld is twofold. Firstly, its direction b gives
the directions to the orthogonal family of lines, and secondly its magnitude j bj is
the measure of the induced frozen-in strains between two solidi ed layers. Likewise,
the weld line, Figure 4.6b, is also uniquely determined from the vector eld b. To
test the orthogonal lines search procedure, any analytical vector function b(r) can
be employed instead of more complex frozen-in temperature gradients. The easiest
is to employ the gradient eld of any arbitrary analytical scalar function for this
purpose. This scalar eld could represent the solidi cation time function tsol (r). In
some cases, there exists an analytical expression for the orthogonal lines, which can
be used for comparison.

6.1.1 Orthogonal Lines for an Assumed Solidi cation


Time Function
It is useful to test the line search procedure using the analytical function for which
isolines are similar to realistic solidi cation fronts. One such two-dimensional func-
tion can be expressed in the following form:

tsol (x; y ) = (xm x)n sin( y + Lx2 ): (6.1)
2K
No attempt has been made to physically describe and explain the constants xm ; n; K
and L. The vector eld b is taken to be the gradient of tsol . Both vector components
bx and by can be easily determined analytically by partial di erentiation of tsol (x; y )
with the respect of x and y. The function (6.1) and its gradient eld are evaluated
for a rectangular computational domain 10  5 units. This region is uniformly split
into 50  25 nite volumes. Figure 6.1a shows the \solidi cation lines" for the
case when the following set of constants are substituted into the solidi cation time

126
Chapter 6 6.1. Validation of the Orthogonal Lines Procedure
equation (6.1):
xm = 11; n = 0:25; K = 3:2 and L = 0:

The solidi cation lines are initiated at three boundary surfaces, the left boundary
look like a plane of symmetry and there is a hot region. With the constant L being
zero, the following analytical expression describes the orthogonal lines:
  2 
y (x; X0 ; Y0 ) =
2 K   
arccos cos 2K Y0 exp  (X0 x)(X0 + x 2xm )
:
 8nK 2
(6.2)
Here, X0 and Y0 are the coordinates of the starting point on the orthogonal line.
They correspond to the coordinates of cell centres during the numerical search pro-
cedure.
Two numerical search methods are analysed and compared with the analytical so-
lution: the rst-order search, described in Figure 5.7, and the second-order search
from Figure 5.8. Results for both procedures are presented in Figures 6.1b and 6.1c
respectively. Here, only the longest orthogonal lines are plotted. The orthogonal
lines are chosen to start in cell centres of most CVs next to the boundary. In the
rst-order accurate method errors accumulate as the orthogonal lines approach the
weld line. Numerically determined points on the orthogonal lines start slightly to
deviate from the analytical solution as the line gets longer, as shown in the weld
line region in Figure 6.1b. The second-order accurate method produces the solu-
tion which hardly di ers from the analytical solutions (Figure 6.1c). Therefore,
the second-order accurate method will be used for obtaining the residual temper-
ature eld Tres hereafter. The computational time required for the calculation of
orthogonal lines is much lower then the time needed to solve calculating the energy
and momentum equations. In both examples, more then 100 orthogonal lines are
determined within a few seconds (Pentium 166MHz).

127
Chapter 6 6.1. Validation of the Orthogonal Lines Procedure

Figure 6.1: Solidi cation fronts and orthogonal lines for the assumed solidi cation
time function: a) solidi cation lines, b) orthogonal lines { the rst-order accurate line
search method, c) orthogonal lines { the second-order accurate line search method.
128
Chapter 6 6.1. Validation of the Orthogonal Lines Procedure
6.1.2 Elliptic Orthogonal Lines
The nal validation case for the orthogonal lines algorithm is chosen to produce
closed, elliptic-shaped orthogonal lines. They are expressed mathematically as:
(x a)2 + (y b)2 = const: (6.3)
a2 b2
The direction eld b of the ellipse (6.3) is normal to its gradient vector, so we have:
b (x; y ) =
2(y b) and b (x; y) = 2(x a) : (6.4)
x y
b2 a2
The computational domain is again rectangular, 10  5 units. This region is divided
into 50  25 nite volumes. The mesh spacing used here is chosen to be non-uniform
along the longer side of the rectangular domain. The rst 25 cell columns are made
50% wider then the remaining 25 cell columns. In Figure 6.2 the orthogonal line
from a single cell centre is determined using various methods. The constants a and
b from (6.3) are set to a = 5 and b = 2:5. The rst-order accurate line search

Figure 6.2: Elliptic orthogonal line.


129
Chapter 6 6.2. Validation of the \ln-cos" Residual Temperature/Stress Solution
method produces the helicoidal trajectory instead of the closed elliptic curve. The
second-order accurate numerical method produces an ideal solution and not even the
non-uniform mesh detoured this line from the analytical solution curve. In Figure
6.2 the orthogonal line for the second-order accurate solution has made several full
loops around the ellipse before it is terminated when the maximum of 1000 points
are determined. Numerical errors are thus considered to be negligible.

6.2 Validation of the \ln-cos" Residual


Temperature/Stress Solution
The one-dimensional transient temperature solution for the case where an in nite
plate cools down symmetrically by applying convective cooling to its boundary sur-
faces is described with the in nite series solution, Equation (3.26) derived in Chapter
3. For some special cases of initial temperature pro le T0 , the number of terms in
the series reduces to one. The analytical solution for the residual temperature eld
and consequent residual stresses exists. One such solidi cation case is modeled nu-
merically in this section.
The following solidi cation case is already considered for thermal analytical solu-
tions in Chapter 3 (Figure 3.5). An in nite polymeric plate of thickness 2L = 10
mm, is cooled symmetrically from its two major surfaces, top and bottom. The
applied convective heat transfer coeÆcient is h = 100 W/m2K and material so-
lidi es at 100 ÆC. The plate is free from all external loads. Table 6.1 summarises
the material properties and cooling constants featuring in the analytical solutions
for temperature, Equation (3.41), residual temperature eld, Equation (3.40) and
residual stresses, Equation (2.28).
Although the problem is one-dimensional, the numerical analysis is set up in two
dimensions to overcome the problem of numerically simulating the zero net-force
130
Chapter 6 6.2. Validation of the \ln-cos" Residual Temperature/Stress Solution
 (kg/m3 ) k (W/mK) c (J/kgK)  (m2 /s)
960 0.4 2000 2:083  10 7

E1 (GPa)  (-) (m/mK) Tg (Æ C)


3.0 0.35 100 100

L (mm) h (W/m2 K) H (1/m) 1 (1/m) C1 (1/m) T0 (Æ C) Tc (Æ C)


5 100 250 186.151 311.693 200 0

Table 6.1: Thermal, mechanical and cooling properties for \ln-cos" validation cool-
ing case.
condition acting at the \in nite" boundary. The theoretical in nite lateral dimen-
sion of the plate is replaced in the numerical analysis with a plate of nite width.
However, the width of the plate is chosen to be relatively large compared to its
thickness. This way the transition region from zero tractions, acting at the plate's
in nite boundary, to the one-dimensional residual stress solution is localised to only
a small portion of the plate near the boundary. The remainder of the plate should
exhibit the stress state described by the analytical solution for an in nite plate.
Details of the numerical analysis in the thickness direction y will be given here only,
and the 2-D cooling cases will be analysed later.
Due to the existence of the symmetry plane at y = 0, only the top half of the plate
will be modelled: 0  y  L. This region is uniformly divided into 20 control
volumes. Four monitoring points are chosen, two being placed in the cell centres
nearest to symmetry plane and the cooling surface. The thickness coordinate (y) of
the monitoring points are 0.125 mm, 2.5 mm, 3.75 mm and 4.875 mm, respectively.
Before solving the energy equation, the initial temperature needs to be assigned to
each cell centre according to Equation (3.41) evaluated at time t = 0. The energy
equation is now solved numerically in 1 second time increments. The comparison
between the calculated temperatures and the analytical solution is shown in Figure
6.3. Numerically calculated temperatures t the analytical solution very well.

131
Chapter 6 6.2. Validation of the \ln-cos" Residual Temperature/Stress Solution

Figure 6.3: Temperature histories for the \ln-cos" validation case { comparison of
the nite volume results with the analytical solution, Equation 3.41, with: 2L = 10
mm, h = 100 W/m2K, 1 = 186:151 m 1, C1 = 311:693 m 1.

Figure 6.4: Residual temperature and residual stresses for the \ln-cos" validation
case { comparison of the Finite Volume results with the analytical solutions, Equa-
tions 3.40 and 2.28, respectively with: 1 = 186:151 m 1, T res = 15:951ÆC.

132
Chapter 6 6.2. Validation of the \ln-cos" Residual Temperature/Stress Solution
As soon as the temperature in all cell centres drops below the solidi cation tem-
perature Tg , the recorded temperature gradients at solidi cation are numerically
integrated to obtain the residual temperature eld Tres. Since the cooling is one-
dimensional, the numerical integration is performed along the y-axis. The point
which solidi es last lies in the central plane (y = 0) and the residual temperature
is set to zero at this location. The reaction of the material to the residual tempera-
ture eld gives the solidi cation-induced residual stresses. The distribution of these
quantities throughout the thickness of the plate are illustrated in Figure 6.4.
The nite volume results match the analytical solution entirely. The residual tem-
perature and the residual stress distribution are the parabolic-like curves, as will be
the case with realistic solidi cation cases. The residual temperatures are all positive,
being the highest at the cooling boundaries. The residual stresses are both tensile
and compressive. The material is in the state of permanent compression on the
cooling boundary surfaces, and in the innermost region of the plate the material is
in tension. It is typical for any plate residual stress problem, regardless of the cause
of residual stresses, that the internal compressive stresses are in equilibrium with its
tensile counterpart. For the plate analysed, the area under the stress distribution
curve in the tensile and the compression regions are equal, Figure 6.4. The average
residual temperature value in the plate T res cannot be determined analytically for
this solidi cation case as the closed form solution for integral of type R lncos(y)dy
does not exist. The average value T res required for the analytical residual stress
solution is determined numerically using Mathcad. For the described solidi cation
case it is calculated as:
Z L Z L
T res = 1 Tres(y )dy =
( Tg Tc )
lncos 1 y dy = 15:951ÆC: (6.5)
L L
0 0

The ln-cos validation case illustrates the fact that the Finite Volume methodology
133
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate
produces very accurate results when solving the energy and momentum equations.
The validation case did not require any special numerical treatments to recover
the analytical solution. Idealised and realistic solidi cation cases represent similar
numerical systems. Therefore, realistic solidi cation cases, for which the analytical
solutions might not be available, can be resolved with con dence using the developed
numerical procedure. The frozen-in strain procedure was incorporated into into the
existing nite volume stress analysis program SA2D [81], which is a house developed
program for solving momentum and energy equations for thermo-elastic solids in
2-D.

6.3 Numerical Studies of Residual Stresses in a


Polymeric Plate
Up to now, the proposed algorithm for predicting residual stresses has been val-
idated numerically using the \ln-cos" analytical solutions. In this section we use
the proposed algorithm to predict residual stresses and associated distortions in a
plate-shaped polymeric product cooled in di erent ways.

Figure 6.5: Geometry of the at plate.

The geometry of the at plate that will be analysed is shown in Figure 6.5 and
134
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate
remains the same for all cooling cases. The plate has a rectangular cross-section 100
mm wide and 10 mm thick. In the longitudinal direction the plate is assumed to
be long, reducing the stress analysis to a two-dimensional problem under the plane
strain condition. Symmetry about plane A-A allows only one half of the plate to
be modelled for all cooling cases. Initially, the plate is entirely in the softened state
T > Tg . This presumed stress-free state is the initial condition for residual stress
analysis.
Cooling commences from a uniform temperature of 200 ÆC. In cooling tests a number
of plate surfaces were subjected to convective cooling conditions. When a cooling
medium is applied to the plate surface, the cooling rate is described by the constant
heat transfer coeÆcient of h = 600 W/m2K and the temperature of the cooling

Figure 6.6: Classi cation of the cooling cases (lines superimposed on the plate ge-
ometry correspond to solidi cation lines.
135
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate
medium is set to Tc = 0ÆC. All remaining surfaces are treated as perfectly insulated.
The cooling is applied along the top, side and bottom surfaces. Figure 6.6 classi es
the cooling cases under investigation.
The cases shown in Figure 6.6 re ect the extreme solidi cation cases in plate man-
ufacture, they are either perfectly symmetrical or asymmetrical. In any event, the
plate solidi es non-uniformly, the result of which are unavoidable, permanent, resid-
ual stresses. However, macroscopic distortions of the plate due to residual stresses
do not always reveal the presence of residual stresses. By using numerical predic-
tion methods, we can deduct the complete residual stress and displacement map
from which conclusions can be drawn about the in uence of di erent manufacturing
process parameters.

6.3.1 Symmetrical Cooling Examples


Symmetrical cooling cases are widespread in technical practice. Here, the same
amount of heat is removed from the hot polymeric plate from its two major surfaces.
Economically, this is the most favourable cooling method since the cooling time is
reduced to minimum. If the manufacturing process is symmetrical, it is expected
that all macroscopic properties, including the displacement eld of the nal plate
product, are also symmetrical. If a symmetrical displacement eld due to residual
stresses is superimposed onto an initially at plate, it will result in a at nal plate
product.
The computational domain for symmetrical cooling cases can be reduced further
to just one quarter of the plate cross-section since all boundary conditions are also
symmetrical with respect to the central plane B-B, Figure 6.5. This cross-section is
subdivided into 150  20 nite volumes along respective coordinate axes, x and y.
Two cases are considered: one-dimensional (1-D) and two-dimensional (2-D) cooling.
136
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate
In 2-D we allow cooling of the plate from its minor side surface also. Thermal and
mechanical material properties used in the simulation are given in Table 6.1. The
nite volume results for the 1-D and 2-D cases are presented in Figures 6.7 and 6.8,
respectively.
The conditions in the vicinity of the symmetry plane A-A are almost identical for
both cooling methods; thus identical distribution of dependent variables is expected
in this region. The side cooling in the 2-D case only in uenced the results locally
and the heat transfer conditions in most of the plate are one-dimensional, where
the temperature histories can be calculated both analytically and numerically. The
comparison of the temperatures at four monitoring points at the plane A-A is shown
in Figure 6.9. In both cases, the solidi cation has started soon after the application
of the cooling medium. Solidi cation fronts represented by isochrones tsol = const
move from cooling boundary surfaces towards the hot interior, Figures 6.7a, 6.8a.
In the 1-D case, the hottest region covers the complete middle surface B-B, while
in the 2-D case it is at the intersection of planes A-A and B-B. Orthogonal lines
stream to these regions as illustrated in Figures 6.7b and 6.8b.
The isolines of the residual temperature eld are plotted in Figures 6.7c and 6.8c.
They are maximum where cooling was the most intense and converge to zero as
they approach the hot region of the domain. The nal residual stress distribution is
the result of the internal equilibrium between the volumetric deformations caused
by the residual temperature eld and the elastic response of the material. Points
with the highest value of Tres normally end in compression. Normal stresses xx
are plotted in Figures 6.7d and 6.8d. Compressive stresses are induced in the outer
layers and are highest on the boundary surfaces. Large surface compressive stresses
can have a twofold e ect on the material. They have bene cial e ect when cracking
of the material is a concern since they elevate the opening stresses required to open

137
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate

e)

Figure 6.7: Results of the FV calculation for the 1-D symmetrical cooling case: a)
solidi cation time isochrones, b) orthogonal lines, c) residual temperature eld, d)
normal residual stress component xx , e) deformed geometry after cooling.

138
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate

e)

Figure 6.8: Results of the FV calculation for the 2-D symmetrical cooling case: a)
solidi cation time isochrones, b) orthogonal lines, c) residual temperature eld, d)
normal residual stress component xx , e) deformed geometry after cooling.

139
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate
the crack. At the same time, they promote dimensional instability of the plate.
They are the driving force for creep of the material and when surface material
is removed in any way, the residual stresses are released from the surface, thus
deforming the remaining material. All this can easily disturb the desired through-
thickness symmetry of the plate.
Deformed shapes for symmetrical cooling cases are drawn in Figures 6.7e and 6.8e.
With perfectly symmetrical cooling conditions plates are at after cooling, apart
from the edge region. Intuitively, this may not have been expected for the 1-D
solidi cation case. However, with no deformation resistance on the edge free surface,
the material deforms to satisfy the zero traction condition. In the edge region, the
centre of the plate is no longer being pulled by the surrounding material and the
near-surface layers are not being pushed by the surrounding material. The most
noticeable di erence in deformed shapes is the increased dilatation of the side surface
in the case of 2-D cooling when compared to that of 1-D cooling.
The through-thickness distribution of the residual temperature eld and residual
stresses in plane A-A are plotted in Figure 6.10. The nite volume solutions are
compared with the approximate analytical solution based on the ln-cos approxima-
tion for the residual temperature eld. The rst root of the eigen-equation (3.14)
for this cooling case equals 1 = 277:564 m 1 and the corresponding mean residual
temperature value is T res = 42:852ÆC. Although the FV and the ln-cos solution
describe di erent cooling cases, the results compare well. This is due to the similar
thermal conditions at the solidi cation temperature Tg and the ln-cos approximation
can give useful information about the residual stresses.
The calculated residual stress distribution for the symmetrical cooling cases are
usually very close to the 2nd order parabola. With a perfect parabolic residual stress
distribution it can be shown that the ratio of the maximum normal compressive vs.
140
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate

Figure 6.9: Temperature histories at the vertical symmetry A-A for symmetrical
cooling cases { comparison of the nite volume results with analytical solution,
Equation 3.28.

Figure 6.10: Residual temperature and normal residual stress component xx at the
vertical symmetry A-A for symmetrical cooling cases { comparison of nite volume
results with approximate analytical solutions, Equations 3.40 and 2.28, respectively
with: 1 = 277:564 m 1, T res = 42:852ÆC.

141
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate
tensile stress magnitude equals:
max(compressive stress magnitude) = xx (L) = 2:
max(tensile stress magnitude)  (0)
xx

In the symmetrical cooling cases the residual stresses can easily reach the level of
tens of MPa. If the polymeric plate is used as a load carrying engineering component,
the residual stresses need to be included in the design of the component.

6.3.2 Asymmetrical Cooling Examples


In asymmetrical cooling cases the heat is mainly removed from the top surface,
while the bottom surface is made adiabatic. Here we shall consider a case with
the cooling conditions and the cooling medium properties being the same as for
symmetrical cases (h = 600 W/m2K, T0 = 200 ÆC, Tc = 0 Æ C). The computational
domain is one half of the plate cross-section due to symmetry plane A-A. This
domain is subdivided into 150  40 nite volumes. Again, cases with and without
side cooling are considered, referred to as the 2-D and 1-D cases, respectively. The
numerical results are presented in Figures 6.11 and 6.12.
With only one major cooling surface, it takes non-proportionally longer for the
temperature in the whole plate to drop below the softening temperature Tg . The last
point solidi es after approx. 206 seconds, compared to 58 seconds for symmetrical
cases. The cooling histories at four monitoring points, all lying in the symmetry
plane A-A, are plotted in Figure 6.13. FV cooling results compare very well with
the analytical solution.
The bottom surface is the hottest region of the domain and the solidi cation fronts
terminate at it, Figures 6.11a, 6.12a. The orthogonal family of lines are all parallel
to each other for the 1-D case, and for the 2-D case the orthogonal lines clearly
142
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate

e)

Figure 6.11: Results of the FV calculation for the 1-D asymmetrical cooling case:
a) solidi cation time isochrones, b) orthogonal lines, c) residual temperature eld,
d) normal residual stress component xx , e) deformed geometry after cooling.
143
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate

e)

Figure 6.12: Results of the FV calculation for the 2-D asymmetrical cooling case:
a) solidi cation time isochrones, b) orthogonal lines, c) residual temperature eld,
d) normal residual stress component xx , e) deformed geometry after cooling.
144
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate
identify the zone of in uence of the side cooling/solidi cation, Figures 6.11b, 6.12b.
The asymmetrical solidi cation is most clearly re ected in the residual tempera-
ture elds, Figures 6.11c, 6.12c. The residual temperatures on the bottom surface
are zero throughout in the 1-D case, and in the 2-D case they rise due to side
cooling. The contours of the xx are presented in Figures 6.11d and 6.12d. The
residual stresses are no longer symmetrical. Compressive stresses exist on both
major boundary surfaces, having the largest magnitude at the top surface, which
cools most rapidly. Regions with the highest residual temperatures are again in
the compression but, unlike in the symmetrical cases, stresses in points with the
lowest residual temperatures are no longer tensile and it is the middle region of
the plate which exhibits tensile stresses. The distribution of residual temperatures
and residual stresses at the symmetry cross-section A-A are given in Figure 6.14.
The calculated asymmetrical stress distribution must comply with the equilibrium
conditions for a non-constrained plate: the net-force and net-moment arising from
residual stresses should equal to zero. Quick qualitative checks of these conditions
can be performed. The zero net-force condition is satis ed if the areas under the
compressive and tensile parts of the residual stress curve are equal. For the zero
net-moment condition, the equivalent concentrated forces can be placed in the cen-
troid of each compressive and tensile parts of the stress distribution curve and the
moments that these forces produce around any point must cancel each other.
The deformed shapes due to residual stresses are shown in Figures 6.11e and 6.12e.
For clarity, deformations are magni ed by a factor of 5. The presence of residual
stresses in the asymmetrical cooling cases is apparent. The plate bends in order
to minimise the strain energy caused by the frozen-in strains. This deformation
reduces to a great extent the residual stresses in the plate when compared with
cases where this deformation is mechanically prevented. The lower level of residual
stresses are bene cial, but its curvature is highly undesirable. The plate always
145
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate

Figure 6.13: Temperature histories at the vertical symmetry plane A-A for asym-
metrical cooling cases { comparison of the nite volume results against the analytical
solutions, Equation 3.28.

Figure 6.14: Residual temperature and normal residual stress component xx at
the vertical symmetry plane A-A for the asymmetrical cooling cases { comparison
of the nite volume results against the approximate analytical solutions, Equation
3.40 with 1 = 147:292 m 1 (analytical residual stress solution will be discussed in
Section 6.3.3).

146
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate
deforms in the \frown" fashion if the cooling medium is applied to the top surface.
The edge e ect is again contained locally. The most noticeable di erence between
the deformed geometries in the 1-D and 2-D asymmetrical cases is the increased
dilatation of the side surfaces when side cooling is applied. Generally, the material
\swells" in regions of intensive cooling. The maximum warp encountered in the
1-D asymmetrical cooling case is Æmax =2.806 mm and the slightly higher warp of
Æmax =2.898 mm in the 2-D case. However, the radius of curvature for both cases
is the same. Asymmetrical solidi cation case results are discussed further in the
following section.

6.3.3 Connection between Symmetrical and Asymmetrical


Residual Stress Pro les
The residual temperature and residual stress pro les have similar shapes for symmet-
rical solidi cation cases as illustrated in Figure 6.10. In asymmetrical solidi cation
cases this similarity disappears, Figure 6.14. However, it will be demonstrated that
the two solutions are linked. Asymmetrical residual stress solutions are contained
within the much simpler symmetrical case results, and can be recovered from there.
Let us assume two plates, P1 and P2, undergoing the same solidi cation processes
and hence have identical residual temperature pro les. Both plates are cooled only
from the top surface. The rst plate P1 is allowed to deform freely due to accumu-
lated frozen-in strains and the second plate P2 is held \ at" by constraining it at
the bottom surface. The former case is equivalent to the symmetrical cooling case
of the plate of a double thickness with only the top half considered. If this plate
is cut at its central plane, the internal stresses that have kept it at are released
and both halves of the plate curve. The dimensions, residual temperatures, residual
stresses and resulting curvatures now match those of the plate P1. The aim here is
to calculate the asymmetrical residual stress pro le for the curved plate P1 by using
147
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate
the symmetrical residual stress pro le determined for the at plate P2.
When the plate of double thickness is cut into two equal parts, the top and bottom
halves no longer physically interact. The mechanical in uence of the bottom half to
the top half can be replaced by the resultant force and the resultant moment in the
dividing cross-section. The force and moment have the e ect of keeping both halves
of the plate at, before the cut is made. For convenience, the resultant force and
moment are moved to the centroid point P of the top half, located at yP = +L=2.
The force and moment are calculated from:
Z+L
Resultant force: FP = res (y ) dy (6.6)
0
Z+L  
L
Resultant moment: MP = res (y ) y
2 dy: (6.7)
0
The resultant force FP in half of the thick plate equals zero (zero net-force boundary
condition). Therefore, two halves of the plate in uence each other only via the
resultant moment MP. The bending stresses due to internal moment MP are already
incorporated in the symmetrical residual stress distribution. When the moment MP
is removed, bending stresses relax and the plate curves. The residual stresses that
remain in the plate result stresses for the asymmetrical cooling case (Figure 6.15).
To quantify the resulting moment and the corresponding bending stresses, we can
substitute the general solution for the symmetrical residual stresses, Equation 2.28,
into the expression (6.7):
Z+L  
E1   L
MP = Tres(y )
0
1  T res y
2 dy:

or 2 3
Z+L
E1 4 L 2
MP = Tres(y ) y dy 5 : (6.8)
1  T res 2 0

148
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate

Figure 6.15: Decomposition of the symmetrical residual stresses into the bending
and asymmetrical pro les.
The bending stresses are linear and are deducted from the conventional bending
stress expression:
 
M L
B (y ) = P y (6.9)
I 2 ;
where I is the moment of inertia for the cross-section of the plate in which bending
stresses are analysed:
Z+L 2
L L3
I= y dy = (6.10)
0
2 12 :
The residual temperature eld and its functions featuring in Equation (6.8) are
usually determined numerically, even for the analytical ln-cos residual temperature
solution these terms cannot be evaluated analytically. The residual stress pro le for
the asymmetrical case is nally determined from:

res (y )
= res(y) symm

B (y ): (6.11)
asymm

The results presented in Figure 6.15 quantitatively correspond to the asymmetrical


cooling case from the previous section. The simpli ed analytical case is described
by the ln-cos solution with the following cooling-dependent constants: 1 = 147:292
149
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate
m 1 and T res = 51:828ÆC. The integral term from Equation (6.8) for this case is
4:072  10 3 m2K and the resultant bending moment equals MP = 683:361 Nm/m.
When this (negative) bending moment is released from the plate, the plate P1
deforms in a convex manner (\frown" shape).
The comparison between the FV asymmetrical residual stress results and the ap-
proximate analytical solution based on the ln-cos approximation has already been
given in Figure 6.14 and a good agreement exists. However, if we use the FV residual
temperature results to calculate the resulting bending moment and then use it to
determine the asymmetrical residual stresses, Equation (6.11), the asymmetrical FV
results are retrieved completely (T res = 49:721ÆC and MP = 645:812 Nm/m). This
con rms the analytical mechanism behind the asymmetrical residual stress pro les.
It still remains to derive the analytical expressions for the radius of curvature and
the corresponding maximum warp due to the applied asymmetrical cooling. The link
between the stresses and strains is given by the constitutive law for the material.
In Chapter 5, two special forms of the Hooke's law were derived for the in nite
plate. These are the rotationally symmetrical and the plane strain forms, given
by expressions (5.7) and (5.9). The strain function is assumed linear and of the
following form:
 
1
"(y ) = "0 + y
L
(6.12)
 2 :
Further, the Hooke's law is substituted into the boundary condition expressions for
zero net-force and zero net-moment and the unknowns "0 and  are thus derived.
For the rotationally symmetrical cases the radius of curvature  becomes:
2 3
Z+L
1 = 12 4T L2
Tres (y ) y dy 5 =
MP
(1  ); (6.13)
 L 3 res
2 0
E1 I

150
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate
and for the plane strain cases the equivalent expression is:
2 3
Z+L
1 = (1+ ) 12 4T L 2
Tres (y ) y dy 5 =
MP
(1  2 ): (6.14)
 L3 res
2 0
E1 I

The radius of curvature  describes the deformed shape of the plate. After cooling,
the plate is usually not curved substantially. Since the radius of curvature is usually
much larger then the width of the plate, the deformed shape of the plate can be
approximated with the 2nd order parabolla to a suÆcient accuracy. The deformed
shape of the plate is simply expressed by:
Æ (x) =
1 x2 ; (6.15)
2
where x is the coordinate along the width of the plate, measured from the vertical
symmetry plane A-A. Equation (6.15) represents the semi-analytical expression for
the plate warpage. The maximum warp Æmax is calculated for the point on the edge
of the plate. The semi-analytical results for the asymmetrical cooling cases described
in the previous section are:
 = 0:441 m and Æmax = Æ (0:050) = 2:833 mm:
The calculated maximum warp is within 2% of the FV results.

6.3.4 Curved Plate Subjected to External Bending


When external load is applied to a linear elastic plate which contains residual
stresses, the resultant stresses in it are determined by the stress superposition prin-
ciple [56]. Sometimes the resultant stresses can lead to somewhat unexpected stress
patterns. In this Section we shall brie y demonstrate one such example.
A plate under examination has been cooled asymmetrically from the top surface.
Due to asymmetrical residual stresses in it, the plate is curved in a \frown" shape.
151
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate

Figure 6.16: Curved plate subjected to external bending moments: a) attening of


the plate by negative external moment, b) curving the plate by positive external
moment (the shapes of non-constrained plates are plotted dashed).
A negative external bending moment M is now applied in an attempt to atten the
plate, Figure 6.16a. The existing compressive residual stresses on the top surface
are further magni ed upon the application of the bending moment. At the same
time, the tensile region has moved from the plate interior to the bottom surface.
If the external moment is further increased the plate may potentially crack on the
bottom surface.
The situation is now reversed when a positive external moment of the same mag-
nitude M is supplied to the plate, Figure 6.16b. The external moment further
increases the curvature of the plate. However, the stress state in this case is more
favourable since the compressive stresses are preserved on both plate surfaces. It is
less likely that this plate will fail by cracking.
152
Chapter 6 6.3. Numerical Studies of Residual Stresses in a Polymeric Plate
This simple example demonstrates that residual stresses can have in uence on the
behaviour of a polymeric product. With the presence of residual stresses, the ex-
pected and conventional behaviour of the component may \surprisingly" change.

153
Chapter 7
Experimental Validation of the
Model

The frozen-in strain model can be used to predict residual stresses in polymeric ma-
terials after thermal processing. Our aim here is to experimentally and numerically
examine the residual stresses and associated distortions for a speci c polymeric ma-
terial. The material of interest is lled polymethyl methacrylate(PMMA) produced
in plate form by a casting process. It is an amorphous glassy polymer with good
mechanical, thermal and aesthetic properties and is used as a replacement for more
expensive materials, such as marble and ceramics. A laboratory technique indicative
of the actual casting process is developed to experimentally determine the response
of the material to applied cooling conditions. This response is later compared with
the results of numerical model.

7.1 Description of the Industrial Casting Process


Casting is a widely used process in the manufacture of polymeric plates (Figure 7.1).
A typical casting process is performed in several stages. During the process the
material transforms from a syrup into the solid state. At the beginning all chemical
components of the material are mixed together to prepare a highly viscous slurry.
154
Chapter 7 7.1. Description of the Industrial Casting Process

Figure 7.1: Schematic diagram of a typical continuous casting production line.

155
Chapter 7 7.1. Description of the Industrial Casting Process
The main ingredients of the slurry include monomer, llers, pigments, initiators,
cross-linking agent, etc. The slurry is then cast onto a moving at metal belt and is
spread uniformly on the belt to achieve the desired dimensions of the future plate.
The prepared slurry is at or near room temperature.

7.1.1 Polymerisation Process


The presence of initiators in the slurry triggers polymerisation. Acrylic polymerisa-
tion is a highly exothermic process: every time a double carbon bond of a monomer
molecule is broken, a xed quantity of heat is released, thus increasing the tempera-
ture of the slurry [82]. The polymerisation kinetics is faster at higher temperatures,
causing more monomer molecules to decompose leading to auto-acceleration of the
polymerisation system. The auto-acceleration becomes apparent when the temper-
ature of the slurry surges [83]. Soon afterwards, the temperature peaks and reaches
the level typically higher than 100ÆC. This temperature maximum is known as the
Trommsdor peak and indicates the end of the process of monomer decomposition
(no additional heat production, no further increase of temperature). At that stage
the polymerisation process is well under way. The slurry becomes more viscous as
polymer molecules grow and the sheet gradually gains its mechanical sti ness. The
preservation of the heat released by polymerisation in the slurry is advantageous to
the growth of polymeric chains. Eventually, the viscosity of the slurry becomes too
high for any further chemical reaction to take place and the material is chemically
structured. Now the material can be characterised as a low sti ness, rubbery poly-
meric sheet. It is sti enough to be withdrawn from the polymerisation belt and is
guided via sets of rollers into the nishing stage, the cooling zone.
The chemistry of the polymerisation process is very complex and is not discussed
here. However, it must be stated that the average material properties after the man-
ufacture are mainly dependent on the chemical composition and the polymerisation
156
Chapter 7 7.1. Description of the Industrial Casting Process
process. The average molecular weight of the polymer and the level of crosslinking
have the main in uence on the nal mechanical properties of the sheet. For the ma-
terial that will be examined in this study, the mechanical properties are determined
experimentally (Appendix A).

7.1.2 The Material at Elevated Temperatures


After polymerisation, the material is at elevated temperature. For lled PMMA
these temperatures are above the glass transition temperature Tg and the material
is in the softened state. The mechanical sti ness of the material is low and it can
easily be handled, bent, stretched or twisted manually. Above the Tg , polymeric
chain segments are movable and upon the application of any load, they rearrange
into a new con guration characterised by the lower strain energy potential [63]. The
stress relaxation process is accompanied by noticeable distortions of the material.
The behaviour of the material at elevated temperatures is similar to the behaviour of
non-Newtonian uids. The material can only partially resist high deviatoric stresses.
Consequently, internal stresses induced into the material during polymerisation are
low and have a minimal e ect on the nal residual stress eld. The con guration
of the material at this stage is called the initial con guration. It can easily be
in uenced by the manufacturer. On the continuous casting production line the
initial con guration is usually at with the sheet being slightly stretched in the
production direction. The rearrangement of polymeric chains results in very low
stresses in the material while at the initial con guration. This presumed \stress-
free" state and the initial con guration characterise the material as it enters the
cooling zone.
For unconstrained plate the initial con guration is not necessarily at, and can be
determined experimentally by the annealing process. Although low, the internal
157
Chapter 7 7.1. Description of the Industrial Casting Process
stresses existing in the plate at high temperature are high enough to \pull" the
material of very low sti ness back to its original shape achieved upon polymerisa-
tion. This is the reason why the polymeric materials are known as materials with
\memory". Every time unconstrained material is placed at temperatures above Tg ,
the material retrieves its \stress-free" con guration.

7.1.3 The Material in the Cooling Zone


During the cooling stage the material solidi es and cools close to the room tem-
perature at the end of the cooling zone. Due to non-uniform solidi cation, frozen-
in strains are generated in the plate. Symmetrical frozen-in strain patterns are
favourable since the sheet will remain at after cooling.
Convective cooling by cold air on both sides of a plate is usually employed as a
cooling method. Cold air is blown from stationary cooling nozzles onto both sur-
faces of the moving plate. The cooling capability of this method is limited mainly
due to low thermal conductivity of the polymer, making the cooling time long and
consequently requiring a long cooling zone. When the temperature falls suÆciently
throughout the plate, it can be cut to required dimensions, stacked and stored.
Flatness of the plate after cooling is of prime importance to the manufacturer. The
control mechanism on the cooling line ensures atness of the sheet after cooling. If
the shape of the sheet becomes curved in any way, the cooling conditions are adjusted
accordingly. For example, if a sheet exhibits a \smile" deformation, cooling must be
intensi ed on the top surface. This is implemented by increasing the heat transfer
coeÆcient on the top surface, which is easily controlled by changing the volumetric
ow of the cooling air.

158
Chapter 7 7.1. Description of the Industrial Casting Process
7.1.4 Properties of the Cast Sheet
A sheet produced on the production line described previously has relatively uni-
formly distributed residual stresses. Residual stress gradients are not expected in
the longitudinal direction since all the points in this direction are subjected to the
same thermal history. In the transverse direction, the cooling conditions can vary,
depending on the local heat transfer conditions. The pro le of cooling nozzles is
designed to ensure the uniform volume ow of cooling air to achieve uniform heat
transfer. Since the sheet is very long in the longitudinal direction, the stress state in
it can be well represented using two-dimensional approximations. The conditions in
the longitudinal direction can be described as plane strain. All this make the process
on the production line convenient from the point of view of numerical modelling.
Unfortunately, heat transfer conditions on the cooling line cannot be determined
easily or with the great con dence. Heat transfer coeÆcients may vary in the lon-
gitudinal and the transverse direction; the heat ow conditions around the edges of
the plate are complex, the initial temperature pro le as the sheet enters the cooling
zone may not be uniform, etc. The uncertainties listed above are potential sources
of error in numerical simulation. Thermal histories that are employed for validation
purposes of the residual stress model should be authentic, no appreciable error is
allowed on the thermal side of the problem.
Thermal histories are veri ed when experimentally measured temperature signals
match numerically calculated temperatures. There are practical problems involving
measuring temperatures in a sheet on the production line. A number of thermocou-
ples must be embedded in the sheet in the earliest stages of the process. Thermocou-
ples are moving with the sheet and signals are continuously captured and recorded.
Although the measured data gives useful information about the polymerisation and
cooling processes, such temperature measurement is very delicate to perform and
159
Chapter 7 7.2. Description of the Cooling Experiment
is very costly. Furthermore, it is not acceptable to shut the production line for
measurement purposes. Alternative methods are needed to validate the numerical
residual stress model and one such lab technique indicative of the actual casting
process is described in the following section.

7.2 Description of the Cooling Experiment


The cooling process from the production line cannot be easily replicated in the
laboratory conditions due to the size of the cooling equipment. Instead, a smaller
laboratory cooling rig is developed to examine the in uence of cooling conditions to
the deformation behaviour of the polymeric plate, Figure 7.2.

7.2.1 Cooling Setup


In the experiment, a preheated polymeric plate sample is subjected to a controlled
thermal cycle. A plate specimen is placed in an oven, heated to a temperature above
Tg and held there until heated throughout. The heat transfer mechanism in the oven
is due to natural convection and radiation and it typically takes 1-2 hours for the
temperature in the plate to stabilise. Afterwards, the hot plate is quickly taken out
of the oven and is placed into the cooling system shown in Figure 7.2. Throughout
the test the plate was resting on the base tray and was moved together with it.
While the plate is moved, it is not handled directly mechanically in any way, since
any applied load would deform it.
The cooling system comprises of a set of cooling nozzles from which the ambient
air is blown vertically onto the plate, Figure 7.3a. Only the top surface of the plate
is cooled; the remaining surfaces are insulated by glass wool blankets to achieve
near-adiabatic conditions. Such highly asymmetrical cooling induces asymmetrical

160
Chapter 7 7.2. Description of the Cooling Experiment

Figure 7.2: Schematics of the cooling rig used for experiments.


frozen-in strains and it is expected that the plate will bend in both directions,
resulting in a \dome" shaped deformed pro le illustrated in Figure 7.3b.

7.2.2 Temperature Measurements


Throughout the experiment, the temperature signals from 15 thermocouples at var-
ious locations on both plate surfaces and within the plate are captured every sec-
ond using the Supreme Mini POD data acquisition system [84], convenient for a
multi-channel thermocouple measurements. The system automatically handles the
speci cs of thermocouple measurements, e.g. calibration curves for thermocouple
type employed and ice point compensation [85]. The measured temperatures are
161
Chapter 7 7.2. Description of the Cooling Experiment

Figure 7.3: Polymeric plate subjected to cooling from ve cooling nozzles: a) cooling
schematics, b) expected dome-shaped plate pro le after cooling.
converted into digital signals and saved to a computer disk.
K type thermocouples (Ni-Cr-Ni) are used for the experiments. Each thermocouple
is attached to the plate by placing it inside a small diameter hole drilled from
the bottom plate surface. The thermocouple wires are glued within the hole. The
thermocouple tip touches the plate material at the bottom of the hole, which de nes
the monitoring location. For the top surface temperatures the holes are drilled all
the way through the plate thickness (Figure 7.3, detail A). Before the thermocouples
162
Chapter 7 7.2. Description of the Cooling Experiment
are attached to the plate, they are calibrated for the melting and boiling points of
water at atmospheric pressure.

7.2.3 Heat Transfer Conditions


Heat transfer conditions on the cooling plate surface are complex. For any system
with forced convection the local Nusselt number is almost constant for a wide range
of surface temperatures. The Nusselt number, Nu, is a non-dimensional representa-
tion for a heat transfer coeÆcient. It contains all speci cs of the cooling system, and
is dominantly dependent on the uid ow conditions, as described by the Reynolds
number (Re) and the properties of the cooling medium characterised by the Prandtl
number (P r). The correlations for the Nusselt number are usually expressions of a
type Nu = Nu(Re; P r) [59]. Special attention should be given when selecting the
right correlation due to many limitations and assumptions used in deriving them.
They are applicable to the extent to which the real cooling system is represented
by the conditions incorporated in the correlation. The cooling conditions for a at
plate can be described by the correlations for impinging jets of gas, blown from the
array of the slot nozzles [86]. A qualitative heat transfer coeÆcient map is presented
in Figure 7.4. The cooling is most intense below the nozzles, whereas between the
nozzles the cooling rates are much lower. In this con guration, the gap between the
cooling nozzles and the plate surface has a signi cant in uence on the local heat
transfer conditions. This cooling gap can be controlled by placing the hot plate to
an adjustable table, Figure 7.2. The alternative to using empirical correlations is
a ow simulation of the cooling medium surrounding the plate. However, resolving
the ow around the plate is beyond the scope of this study.
The purpose of thermocouple measurements is to indirectly determine the heat
transfer coeÆcients on the cooling surface. At rst, the heat transfer coeÆcients are

163
Chapter 7 7.2. Description of the Cooling Experiment

Figure 7.4: Variation of local heat transfer coeÆcients for arrays of slot nozzles,
from [86].
assumed for the whole plate. The assumed distribution should re ect the qualitative
heat transfer coeÆcient map from Figure 7.4. Secondly, the assumed heat transfer
coeÆcients are used to carry out the numerical analysis. As a result, temperature
histories are determined for the whole plate, including the thermocouple locations.
All measured temperature signals are compared with the calculated temperatures.
Underestimated local heat transfer coeÆcients result in a higher calculated temper-
ature and need to be increased in subsequent numerical calculations. The iterative
process of adjusting the heat transfer coeÆcients is repeated until the calculated
temperature histories match the measured temperatures at all monitoring locations.
With the thermal history from the experiment well replicated in the numerical anal-
ysis, the residual temperature eld and residual stresses can be determined.

7.2.4 Distortions of the Plate


After the experiment the test plate exhibits substantial distortion due to residual
stresses. The deformed shape of the plate can be measured on a Coordinate Mea-

164
Chapter 7 7.2. Description of the Cooling Experiment
suring Machine (CMM) [87]. This is a computer-controlled measuring system which
determines the coordinates of points on a solid surface using a very sensitive mechan-
ical probe. The mechanical probe touches the surface at a number of pre-de ned
locations and records their locations onto a computer disk. The measured coordinate
map gives a clear picture of the distortion after cooling. The CMM measurement
is a very swift experimental technique as it takes only a few minutes to measure
hundreds of points on the plate surface.

7.2.5 Annealing Pro les


As mentioned previously, it is possible to in uence the initial \stress-free" shape
of the plate at temperatures above the Tg . It is important to determine the initial
shape, as the distortion induced during cooling is superimposed upon it. In our
cooling experiments the specimen plate is left unconstrained in the oven where it
anneals. The annealed pro le could signi cantly di er from the at pro le, and it
needs to be measured and used as the reference pro le. It would be ideal to measure
this pro le prior to cooling when the plate is at the initial (elevated) temperature.
Unfortunately, it is not possible to use CMM to measure the shape of the plate while
it is in the oven. Therefore, the plate needs to be brought to room temperature
with the annealed pro le preserved. This can be accomplished by subjecting the
plate to an extremely slow cooling regime. This minimises temperature gradients
at solidi cation and the frozen-in strains induced in the plate are practically zero.
After slow cooling the plate shape is measured by CMM in conventional manner. A
number of plates can be annealed during the same annealing cycle. For the purposes
of the cooling experiments, the oven temperature has been gradually reduced from
120ÆC to 90ÆC over 62 hours. For all examined plates the measured annealed
pro les were saddle-like surfaces.

165
Chapter 7 7.3. Cooling Experiments with Filled PMMA Plates
7.3 Cooling Experiments with Filled PMMA Plates
A number of experiments have been performed to experimentally validate the pro-
posed residual stress model. In the experiments, the plate specimens, with dimen-
sions 25412712:7 mm, have been used (Figure 7.3a). The origin of the coordinate
system is placed in the middle of the bottom plate surface (-127  x  127 mm,
-63.5  y  63.5 mm, 0  z  12.7 mm). All test plates were subjected to asym-
metrical cooling on the cooling rig, Figure 7.2. Cooling conditions varied from test
to test, primarily by changing the oven temperature, the number of cooling nozzles
and the cooling gap. There are maximum of ve cooling nozzles equally placed along
the 254 mm cooling distance.
In the rst cooling example we have used all ve nozzles to cool the test plate. The
plate was preheated in the oven to 142.5ÆC. Throughout the heating and cooling
cycles, temperatures were monitored at 7 characteristic cross-sections, measuring top
and bottom surface temperatures. Owing to symmetry about the central cooling
nozzle, Figure 7.3a, the thermocouples were placed only in the right half of the
plate, x  0 mm. Furthermore, due to uniform cooling conditions in the width
direction y, there was no need to distribute thermocouples along the width of the
plate. All monitoring locations were chosen along the central cross-section, y = 0
mm. Their positions are: (a) 0, (b) 10, (c) 31.75, (d) 63.5, (e) 73.5, (f) 95.25, (g) 125
mm along the x-axis. Cross-sections (a), (d) and (g) are facing the cooling nozzles,
where the temperatures are expected to drop fastest. Cross-sections (c) and (f)
are located exactly between the neighbouring cooling nozzles where cooling is less
intense. Thermocouple readings for all cross-sections are shown in Figure 7.5.

166
Chapter 7 7.3. Cooling Experiments with Filled PMMA Plates

Figure 7.5: Comparison of temperature histories at characteristic locations on the


plate after appropriate heat transfer coeÆcient map is determined for numerical
analysis.
167
Chapter 7 7.3. Cooling Experiments with Filled PMMA Plates
7.3.1 Thermocouple Measurement Results
Top surface temperatures drop very quickly below the solidi cation temperature
(Tg = 105ÆC) in the nozzle cross-sections, approx. 1 minute from the start of cooling,
Figure 7.5(a). The most rapid cooling was recorded in the edge cross-section (g).
Here, the plate is additionally cooled from the side surface. In the area between the
nozzles, (c) and (f), the solidi cation front has initiated within 2 minutes after the
start of cooling. Although the bottom surface of the plate was insulated, some heat
was also conducted from there. All temperature history curves from the bottom
surface show immediate temperature drop. If the bottom surface was perfectly
insulated, it would take some time before the temperature there started dropping.
It is worth noting the similarity between the measured temperatures in the cross-
sections subjected to similar cooling conditions, (a) and (d), (b) and (e), (c) and
(f). This implies that the heat transfer conditions were close to one-dimensional for
most of the plate.

7.3.2 Finite Volume Heat Transfer Analysis


The measured temperatures are used to de ne the boundary conditions for the
energy conservation equation in the Finite Volume analysis. The FV analysis was
set up as two-dimensional, and only one half of the central plate cross-section was
modelled (0  x  127 mm, 0  z  12.7 mm). This domain was discretised with
100  15 cells. In the analysis the following boundary conditions were used:
 Left boundary { symmetry plane (adiabatic surface),
 Top boundary { spatially distributed heat transfer coeÆcient h = h(x),
 Right boundary { constant heat transfer coeÆcient h = 109 W/m2K,
 Bottom boundary { spatially distributed heat ux q = q(x).
The thermal material properties are as follows:
168
Chapter 7 7.3. Cooling Experiments with Filled PMMA Plates
 Thermal conductivity, k = 0:6 W/mK,
 Speci c heat capacity, c = c(T ) (Appendix B),
 Density,  = 1754 kg/m3.
After several numerical trials, the heat transfer coeÆcient map was determined,
Figure 7.6, such that the heat transfer conditions during the experiment are well
represented. All numerically calculated temperature histories tted well the mea-
sured temperatures (Figure 7.5). The distribution of the heat transfer coeÆcient
along the plate from Figure 7.6 is qualitatively similar to the theoretical distribu-
tion from Figure 7.4. The heat transfer coeÆcient is almost twice as high below
the nozzle comparing to mid-nozzle areas. On the bottom surface, the heat ux of
approx. q = 1800 W/m2 was found to describe the conditions in the experiment
well.

Figure 7.6: Heat transfer coeÆcient distribution on plate cooling surface. Cross-
sections (a), (b) ...(g) are de ned in Section 7.3

As a result of the numerical analysis, temperature histories are obtained for all
points in the plate cross-section. This uniquely de nes the solidi cation process
and the frozen-in strains induced. Figure 7.7a shows the predicted progress of the
169
Chapter 7 7.3. Cooling Experiments with Filled PMMA Plates

Figure 7.7: Results of the FV frozen-in strain calculation for 5 nozzle cooling exper-
iment: a) solidi cation lines, b) orthogonal lines, c) residual temperature eld.
solidi cation fronts during the test. Since the cooling is most intensive below the
cooling nozzles, solidi cation began there and spreads towards the warmer interior.
The second solidi cation front started from the bottom surface and both fronts met
in the warm region situated within the bottom half of the plate. The points that
solidify last are in the hot regions, clearly visible in the mid nozzle regions. The
integration lines are presented in Figure 7.7b. They show that the heat during
solidi cation has mainly been conducted in the thickness direction, apart from the
localised edge region. Integration lines stream towards the weld line, the region with
the minimal strains induced between the neighbouring layers. The resulting residual
temperature pro le is plotted in Figure 7.7c. It is positive throughout the domain,
being the highest below the cooling nozzles and the lowest in the weld line region.

170
Chapter 7 7.3. Cooling Experiments with Filled PMMA Plates
7.3.3 Finite Volume Residual Stress Analysis
The FV method applied to the momentum conservation equation has been set up
as a three-dimensional problem since none of the two-dimensional approximations,
i.e. plane strain or rotational symmetry, can be used to describe the deformation
state of the cooling plate to suÆcient accuracy. Due to the symmetry about the
central cooling nozzle and across the width, only 1/4 of the plate was modelled. The
geometry was subdivided into 100  20  15 control volumes in x; y and z directions,
respectively. For convenience, the initial plate shape is set to be perfectly at so that
the numerically predicted deformed shape is due solely to cooling-induced residual
stresses. The material properties used in the analysis are:
 Long term Young's modulus at room temperature, E1 = 9:62 GPa,
 Poisson's ratio,  = 0:33,
 CoeÆcient of linear thermal expansion at the softening temperature,
(Tg ) = 90:4m/mK.

The calculations were performed using the nite volume code FOAM [7, 88] and the
results are given in Figure 7.8. The residual temperature pro le from Figure 7.7 was
applied in all cross-sections in the width direction (Figure 7.8a) as input data for the
momentum equation analysis. The resulting normal residual stresses xx and yy are
plotted in Figures 7.8b and 7.8c. Both stress components are compressive on plate
surfaces and tensile in the middle. Similar to the two-dimensional cooling cases,
intensive cooling from the top surface causes the hogging deformations of the plate.
The plate curves in both x and y directions, producing a dome-shaped pro le. The
solution obtained must comply with the equilibrium conditions, i.e. zero net-force
and zero net-moment. As a quick qualitative check of the zero net-force condition,
the amount of compression and tension needs to be equal in the y-z symmetry plane
for the stress component xx, Figure 7.8b. Likewise, the yy compressive and tensile
stresses must cancel each other in the x-z symmetry plane, Figure 7.8c.
171
Chapter 7 7.3. Cooling Experiments with Filled PMMA Plates

a)

b)

c)

Figure 7.8: Results of the 3-D FV residual stress calculation for 5 nozzle cooling ex-
periment: a) applied residual temperature eld, b) normal residual stress component
xx , c) normal residual stress component yy (contours presented on the deformed
geometry). 172
Chapter 7 7.3. Cooling Experiments with Filled PMMA Plates
The stress component xx, Figure 7.8b, is very uniformly distributed throughout the
plate. It is diÆcult to deduce the location of the cooling nozzles from xx contours.
The xx stress concentrations are not visible in the nozzle areas. The expected high
stresses could have relaxed only by increased local deformations in the nozzle areas.
The deformed pro le along the x-axis will therefore have regions of the increased
curvature, a \knee" below each cooling nozzle. Quite opposite, yy contours (Figure
7.8c), show more non-uniform distributions. However, as the cooling was uniform
in the y-direction, a single radius of curvature is suÆcient to describe the deformed
pro le along any line parallel to the y-axis.

7.3.4 Numerical Predictions vs. Experiment


The experimentally measured deformed shapes are shown in Figure 7.9. The nal
deformed shape after cooling, Figure 7.9c, consists of two independent superimposed
pro les: the annealed pro le of Figure 7.9a and the cooling-induced pro le of Figure
7.9b. The annealed and nal pro les were measured by CMM. Their subtraction
results in the experimental cooling-induced pro le, which should correlate with the
numerically predicted deformations.
The annealed pro le is usually a saddle-like surface, believed to be set in the material
during non-uniform polymerisation. Many plates in the past were subjected to
similar cooling conditions, but the nal shapes after cooling were very di erent. By
disregarding the annealed pro le, the nal shape after cooling may not closely re ect
the deformations induced during the cooling process.
The comparison between the numerically predicted cooling-induced deformed shapes
and the measured pro les are presented in Figure 7.10. For clarity, only the deformed
pro les at both central cross-sections, at y = 0 and x = 0 (Figure 7.9b), are extracted
from the three-dimensional results and presented on the graph. They are in very
173
Chapter 7 7.3. Cooling Experiments with Filled PMMA Plates

Figure 7.9: Deformed pro les determined experimentally for 5 nozzle cooling exper-
iment: a) shape after annealing { CMM, b) distortion induced by cooling, c) nal
deformed shape after cooling { CMM.
174
Chapter 7 7.3. Cooling Experiments with Filled PMMA Plates

Figure 7.10: Experimental and numerically predicted deformed pro les at both cen-
tral cross-sections for 5 nozzle cooling experiment (hogging deformations due to
more intensive cooling from top surface).
good agreement. Close examination of the deformed pro le in the x-direction reveals
the location of the cooling nozzles in both the numerical and experimental pro les.
In the y-direction these pro les are smooth. The maximum de ection along the x-
axis is approximately Æ =1.4 mm. Rather high external loads are needed to deform
the plate to that extent which shows how signi cant the cooling-induced residual
stresses may be.
In another experiment, the cooling was performed using only a single, centrally
located nozzle. The plate is cooled from the initial temperature of 150.8ÆC. Only
the nal results are presented here. The experimentally measured shapes for the
175
Chapter 7 7.3. Cooling Experiments with Filled PMMA Plates
single nozzle cooling experiment are shown in Figure 7.11. The nal shape after
cooling, Figure 7.11c, was found to be signi cantly di erent from the equivalent
pro le from the 5 nozzle cooling case, Figure 7.9c. The most noticeable di erence is
the deformation in the width direction (y), where the latter plate exhibited \smile"
deformations. The reason for the discrepancy lies in the di erence in the initial,
annealed pro les.
The annealed pro le in Figure 7.11a is again the saddle-like surface. However, the
orientation of the plate has been chosen to be the opposite to that in the previous
cooling experiment, Figure 7.9a. By subtracting this pro le from the nal deformed
pro le, the deformation induced by cooling is determined, Figure 7.11b. The plate
hogs in both directions, resulting in a dome-like surface. The deformed pro les
in both central cross-sections, as measured and predicted numerically, are given in
Figure 7.12. Below the central nozzle the cooling pro le is substantially curved in
the x-direction and the remaining areas are almost at. The measured and predicted
pro les compare well. From the last two sets of results, we can deduce the general
trend about distortions due to thermally induced residual stresses: a softened, at,
polymeric plate cooled from the top surface will hog on cooling to room temperature.

The cooling experiments represent very sophisticated solidi cation cases and most
of the features observed on the experimentally measured cooling-induced pro les
were predicted with the numerical residual stress model. It is to expect that the
process from the production casting line, with the appropriate thermal conditions,
can be modelled with even greater con dence. Although we have only compared the
numerical and experimental deformed pro les, the total experimental validation of
the numerical model would require the comparisons of the residual stress pro les.
Unfortunately measuring very non-uniform residual stress elds is complex and in-

176
Chapter 7 7.3. Cooling Experiments with Filled PMMA Plates

Figure 7.11: Deformed pro les determined experimentally for 1 nozzle cooling case:
a) shape after annealing { CMM, b) distortion induced by cooling, c) nal deformed
shape after cooling { CMM.
177
Chapter 7 7.3. Cooling Experiments with Filled PMMA Plates

Figure 7.12: Experimental and numerically predicted deformed pro les at both cen-
tral cross-sections for 1 nozzle cooling experiment.
volves many uncertainties in the measuring techniques. Reliable measured stress
pro les for comparison purposes are diÆcult to obtain. A few preliminary mea-
surements using hole drilling methods show residual stress levels close to numerical
predictions [89].
In the following Chapter we shall analyse some aspects of the residual stress mea-
surements and use the nite volume method to illustrate some of their theoretical
and numerical features.

178
Chapter 8
Applications of the Residual
Stress Model

In the previous Chapter, the proposed model for predicting residual stresses was
validated against experimental results. Here in this Chapter, some practical situa-
tions that can be resolved using the numerical model will be illustrated. Theoretical
and numerical aspects of the Layer Removal Method (LRM) are described. The
LRM is a widely used method for measuring residual stresses in sheet materials.
The sheets that are being examined numerically are cooled on single and multiple
cooling lines respectively. In the latter cooling method, the residual stresses in the
sheet are substantially reduced. The Finite Volume simulation of the layer removal
correlates well with the layer removal theory.

8.1 Layer Removal Method


Measuring residual stresses is a complex process. The residual stresses are deter-
mined indirectly via other dependent variables that can be measured more easily, e.g.
strains and curvature. However, the conversion from measured values into residual
stresses is a cumbersome process, and can introduce substantial errors. The source of
errors lies in the experimental precision of the measurements and more importantly,
179
Chapter 8 8.1. Layer Removal Method
to the degree to which the actual sample complies with the assumptions incorpo-
rated into the stress conversion algorithm. Here we shall concentrate on the layer
removal method, a well-established experimental procedure suitable for measuring
residual stresses in the sheet materials [48].

8.1.1 Theoretical Description of the Method


The layer removal method is a destructive method used to determine the through-
thickness distribution of residual stresses in sheet materials. In the method, suc-
cessive uniform layers of the material are removed from the surface of the specimen
plate. The removal of the surface material and consequently stresses in it, disturbs
the internal mechanical equilibrium in plate. The plate responds by redistribut-
ing the internal stresses to achieve the new equilibrium state. This redistribution
of internal stresses is accompanied by a visible distortion of the plate. The plate
curves and its curvature can be measured experimentally, e.g. by using CMM (Sec-
tion 7.2.4). By measuring the curvature of the plate after each layer is removed,
the original residual stress pro le in the section can be deduced from the measured
curvature.
The derived relation between the measured curvature and the residual stresses will
be based on the following assumptions:
 The residual stresses in the plate vary only in the thickness direction (z) and
are constant in all other directions.
 The material is linearly elastic and responses with linear stress distributions
when subjected to bending loads.
 The material is isotropic and homogeneous, i.e. E =const,  =const.

180
Chapter 8 8.1. Layer Removal Method
 The process of machining layers o the plate surface does not induce additional
residual stresses into the material (the temperatures in the cutting region are
below the softening temperature Tg , no local plastic deformations).

Figure 8.1: Geometry of the plate element with the layers to be removed.
The geometry of the plate element subjected to layer removal is shown in Figure 8.1.
The plane z = +z0 represents the top surface of the original plate and during layer
removal this surface shifts to a new location, given by z = z1 . Before any layer is
removed from the surface of the unconstrained plate specimen, the normal residual
stress xx (z) exists in the cross-section of the plate normal to the x-axis. Similarly,
the stress yy (z) exist in the cross-section normal to the y-axis. The stresses in
cross-sections are in the mechanical equilibrium and produce a zero net-forces and
zero net-moments:
Zz0 Zz0
Net-forces: xx (z ) dz = 0 and yy (z ) dz = 0; (8.1)
z0 z0
Zz0 Zz0
Net-moments: xx (z ) z dz = 0 and yy (z ) z dz = 0: (8.2)
z0 z0
If a layer of the material is now removed from the top surface of the plate, the plate
deforms. Its deformation can be prevented by applying external loads, forces and
181
Chapter 8 8.1. Layer Removal Method
moments, equal to the net-forces and the net-moments that existed in the removed
section. Upon their removal, the plate would elongate/contract and curve in both
directions, x and y. It is preferable to determine the required forces/moments from
the measured curvatures, x and y only, since the curvature can be measured
more accurately then the elongation/contractions of the plate. Due to the initial
equilibrium, Equations 8.1 and 8.2, the net-forces/moments in the remainder of the
plate ( z0  z  z1 ) are in the equilibrium with the forces/moments that have kept
the plate undeformed. Therefore, for the cross-sections normal to x-axis they can
be calculated from:
Zz1
Remaining force: Fx (z1 ) = xx (z ) dz; (8.3)
z0
Zz1 h z0 z1 i
Remaining moment: My (z1 ) = xx (z ) z +
2 dz: (8.4)
z 0

In Equation (8.4), the centroid of the cross-section has been moved from the initial
position z = 0 to the new location z = (z0 z1 )=2. The easiest way to derive the
required expression for xx = xx ( x; y ) is by di erentiating Equation (8.4) with
the respect to z1 :
Zz Zz
dMy (z1 )
1 1
d h i d h z0 z1 i
dz1
= dz1
xx (z ) z dz +
dz1 2 xx (z ) dz = (8.5)
0z z 0
Zz1
= z0 +2 z1 xx (z1) 21 xx (z) dz = z0 +2 z1 dFdzx(z1 ) 21 Fx(z1 ):
1
0 z
Rearranging the terms in (8.5) we obtain:
d h Fx (z1 ) i 2 dMy (z1) :
dz1 (z0 + z1 )
= (z0 + z1 )2 dz1 (8.6)
Integration of the last expression in the limits [z1 ; z0 ] yields:
Zz
Fx (z1 ) =
2 My (z1 )
+ 4 (z0 + z1 )
1
My (z )
dz: (8.7)
(z0 + z1 ) z
( z0 + z )3
0

182
Chapter 8 8.1. Layer Removal Method
The zero net-force and the zero net-moment conditions, Fx (z0) = 0 and My (z0 ) = 0
respectively, are incorporated into the last expression. The expression (8.7) relates
the external force and the external moment required to keep the plate subjected to
layer removals at. If it is di erentiated once more with respect to z1 , the residual
stress distribution can be determined [48]:
Z z0
dF (z )
xx (z1 ) = x 1
dz1
= (z + z ) dz + (z + z )2 4 (zM+y (zz))3 dz: (8.8)
2 dMy (z1 ) 2 My (z1 )
0 1 1 0 1 0
z 1

The residual stress distribution xx is a function of the moment My required to keep
the plate at after layers are removed from it. The formula (8.8) consists of three
terms: derivative of the moment, the value of the moment and the integral of it. The
main diÆculty in using it arises from the need to estimate the derivative term, it is
very sensitive to experimental errors. Unfortunately the derivative term is dominant
in the near-surface regions where the calculated stress is prone to increased errors.
On the top surface of the plate (z = z0 ), the residual stress is estimated only from
the derivative term:
 (z ) =
1 dMy (z1 ) : (8.9)
xx 0
z0 dz1 z1 =z0

The residual stress equation (8.9) does not contain any material property constants
and hence can be used for any general material behaviour. However, for linear elastic
material we can expand (8.8) using the plate bending equations [90]. These relate
the applied moment My to the principal curvatures of the plate x and y :
E (z0 + z1 )3
My (z1 ) =
1  2 12 [ x (z1) +  y (z1 )] : (8.10)
Here, E and  are the Young's modulus of the material and the Poisson's ratio
respectively. The positive moment My curves the plate in a \frown" manner along
the x-axis. By convention, this will be considered as a positive curvature with the
\smile" deformation considered negative. The inverse of curvature is the radius
183
Chapter 8 8.1. Layer Removal Method
of curvature of the deformed geometry :
1 and (z ) = 1 (+ve = \frown", -ve = \smile").
x (z1 ) =
x (z1 ) y 1
y (z1 )
(8.11)
Releasing the at plate of a moment My is equivalent to superimposing onto it a
negative moment My , which would result in a \smile" deformation.
By substituting the negative moment My from (8.10) into the residual stress for-
mula (8.8), we obtain the nal layer removal expression for linear elastic materials
[48]:
(
2 d x (z1 ) d y (z1 ) i
E h
xx (z1 ) =
6(1  2 ) (z0 + z1 ) dz1 +  dz1 +
Zz0 )
   
+ 4(z0 + z1) x(z1 ) +  y (z1 ) 2 x (z) +  y (z) dz : (8.12)
z1

In a special case where the plate remains at in the longitudinal direction ( y = 0),
we obtain the simpli ed version of the Equation above:
Zz0
E h d (z ) i
xx (z1 ) = ( z0 + z1 )2 x 1 + 4(z0 + z1 ) (z1) 2 (z) dz :
6(1 )
2 dz1 x
z1
x

(8.13)
This solution will be compared with the nite volume solution in the following
sections.

8.1.2 Curvature Functions


Curvatures x(z1 ) and y (z1) are measured in discrete increments of z1 . The mea-
sured values cannot be used directly in Equation (8.12), as their derivatives and
integrals are also required. The simplest piece wise linear interpolation functions for
x (z1 ) and y (z1 ) between the measured points are not appropriate for Equation

184
Chapter 8 8.1. Layer Removal Method
(8.12), as they would produce discontinuous residual stress pro les due to discontin-
uous rst derivatives at the measured points. Higher order interpolation functions
between the measured points, i.e. cubic spline interpolation, are also not appropri-
ate although their rst derivatives are continuous throughout. Due to experimental
errors, unphysical variations in the rst derivatives (changes between positive and
negative curvature at in ection points) can substantially distort the residual stress
pro les.
To lter out the experimental errors from the measured curvature points, the least
square polynomial t may be used to describe the curvature functions x(z1 ) and
y (z1 ). The order of the tted polynoms must be low (linear, parabolic, cubic) to
achieve smooth residual stress pro les. With the polynomial tting curves employed
to describe the curvature functions x(z1 ) and y (z1 ), the residual stress function
xx (z1 ) is also polynomial of order increased by one. The parabolic residual stress
pro les therefore result in linear changes in curvature during the layer removal ex-
periment.

8.1.3 The Reverse Problem


If the residual stress pro le for the plate is known, it may be useful to solve the
reverse problem and determine the curvature pro les that the layer removal method
would produce. Such curvature functions are direct indicators of the in uence of
residual stresses to the deformation behaviour of the plate. The solution of the
reverse problem can be formulated from the residual stress equation (8.8). After
di erentiating it with the respect to z1 , we obtain a second order di erential equa-
tion:
d 2 My (z1 ) z0 + z1 dxx (z1 )
dz12
= 2 dz1
: (8.14)

185
Chapter 8 8.2. Cooling a Cast Plate in Multiple Cooling Zones
The unique solution for the moment My (z1) can be obtained after specifying two
boundary conditions. The simplest choice here is to express the moment My and its
rst derivative on the top surface z1 = z0 (from the zero net-moment condition and
Equation 8.9 respectively):
dMy (z1 )
dz1

z1 =z0
= z  (z ) and M (z ) = 0:
0 xx 0 y 0 (8.15)

In case of symmetrical residual stress pro les in the plate, expressed in terms of the
residual temperature eld Tres(z1 ) (equation 2.28), the di erential equation (8.14)
becomes:
d 2 My (z1 ) E z0 + z1 dTres (z1 )
dz1
2 = 1  2 dz1
; (8.16)
with the corresponding boundary conditions:
dMy (z1 ) E  
dz1 z =z 1 
1

0
= z0 T res Tres (z0 ) and My (z0 ) = 0: (8.17)
The di erential equations (8.14) and (8.16) are in a form suitable for numerical
di erential equation solvers, i.e. Runge-Kutta algorithms [80]. Once the solution
for My (z1 ) is obtained, the curvature can be calculated via the bending expression
(8.10). All aspects of the layer removal algorithm will be illustrated numerically
using two plates subjected to di erent cooling parameters.

8.2 Cooling a Cast Plate in Multiple Cooling Zones


On conventional casting production lines, the polymerised sheet is cooled by cold
air in the cooling zone. A large temperature di erence between the plate surface
and the cooling air results in the rapid heat transfer very early in the cooling zone.
This increased heat ow is conducted entirely from the plate surface, thus creating
large temperature gradients. The plate solidi es early in the cooling zone when the
186
Chapter 8 8.2. Cooling a Cast Plate in Multiple Cooling Zones
temperature gradients are large. Consequently, the frozen-in strains accumulated in
the material are also substantial. Post-solidi cation cooling to the room temperature
usually takes considerably longer. Slow initial cooling can substantially increase the
time to solidi cation, while the overall cooling time is not increased to a great extent.
With slow initial cooling followed by rapid cooling after the plate is solidi ed (T < Tg
everywhere within the plate), the resulting residual stresses will be much lower, and
they can easily be halved when compared to conventional cooling conditions. The
slower initial cooling can be achieved by using warm air as a cooling medium. Here,
we will numerically model one such multi-cooling zone example.

8.2.1 Finite Volume Cooling Simulation


A long at plate of dimensions 254  12:7 mm is cooled symmetrically in single and
multiple cooling zones. In the single cooling zone a 25ÆC air is employed throughout
the cooling zone. In the multiple cooling zone, the cooling is performed in two stages.
During the initial 150 seconds of cooling, air at 75ÆC is used. After the initial
phase, the air temperature is lowered to 25ÆC. The temperature of the plate after
the polymerisation is assumed to be uniform at 140ÆC. In the numerical analysis
we shall modell only one half of the plate, due to vertical symmetry. This region
is split uniformly into 100  16 control volumes. The sizes of CVs are 4x = 1:270
mm and 4z ' 0:8 mm. Table 8.1 summarises the thermal, mechanical and cooling
properties used in the numerical analysis.
 (kg/m3 ) k (W/mK) c (J/kgK) E1 (GPa)  (-) (m/mK) Tg (Æ C)
1754 0.6 1500 7.62 0.35 90.4 105

z0 (mm) h (W/m2 K) 1 (1/m) T0 (Æ C) TCZ 1 (Æ C) TCZ 2 (Æ C) tCZ 1 (s)


6.35 125 149.398 140 75 25 150

Table 8.1: Thermal, mechanical and cooling properties for multiple cooling zone
nite volume simulations.
187
Chapter 8 8.2. Cooling a Cast Plate in Multiple Cooling Zones
In both cooling examples the temperatures were recorded for cells in the vertical
symmetry plane. Due to symmetrical cooling conditions on the top and bottom
surfaces, only 8 out of 16 traces need to be recorded. The temperature histories
from a single and multiple cooling zones simulations are presented in Figure 8.2

Figure 8.2: Calculated temperature histories for a single and multiple cooling zones
test simulations.
In the multiple cooling zones example, the cooling zone 1 (CZ1) is devoted entirely
to the solidi cation process. The length of it must ensure that the warmest re-
gions of the plate reach the solidi cation temperature Tg at the end of CZ1. This
completes the frozen-in strain generation process and the plate can be subjected to
more intensive cooling afterwards, in the cooling zone 2 (CZ2). Figure 8.2 clearly
illustrates the proportion of times that the plate spends in various cooling zones.
For the cooling parameters employed, the length of CZ1 is approx. 25% of the total
length of the cooling line, and CZ2 covers the remaining 75%. The plate is con-
sidered ready for storage when temperatures throughout drop below 35 Æ C. In the
single cooling zone case the total cooling time is 506 seconds. This time is increased

188
Chapter 8 8.2. Cooling a Cast Plate in Multiple Cooling Zones
to 585 seconds in the multiple cooling zone. This represents the 15% increase in the
total length of the cooling zone. However, the solidi cation time for the slow cooling
case has almost doubled when compared to conventional, fast cooling case. Much
lower frozen-in strains are therefore expected in the multiple cooling zones case.

Figure 8.3: Residual temperature elds for single and multiple cooling zones cases
(ln-cos approximation obtained from Equation 3.40 with 1 = 149:398 m 1).
The residual temperatures for both cooling cases are presented in Figure 8.3. As
expected they are substantially lower in the slow initial cooling case. Additionally,
the ln-cos approximation for the fast initial cooling case is also plotted and compares
well with the nite volume results. Since the heat transfer coeÆcient has been
kept the same for both cooling cases, the rst root 1 of the eigen-equation (3.14)
is also the same ( 1=149.398 m 1). Hence, the ratio of residual temperatures is
proportional to the temperature di erences between the softening temperature Tg
and the cooling temperatures TCZ 1 and TCZ 2:
Tres(z0 ) fast 41:2 Tg TCZ 2 105 25
=
Tres (z0 ) slow 17:8
= 2:31 and =
Tg TCZ 1 105 75
= 2:67: (8.18)

189
Chapter 8 8.3. Finite Volume Layer Removal Simulations
The calculated ratios di er slightly since the residual temperature pro le for the
slow cooling case is distorted and does not follow the ln-cos curve in the central
region of the plate. This is caused by the fact that the plate left CZ1 before it
has solidi ed throughout and the cold air from CZ2 has in uenced (increased) the
solidi cation temperature gradients in the central region. The calculated residual
temperature pro les will be used in the nite volume layer removal simulations in
Section that follows.

8.3 Finite Volume Layer Removal Simulations


The layer removal can be simulated numerically using the previously developed
residual stress algorithm. The material responds with residual stresses and dis-
tortions to any applied residual temperature eld. If the material does not soften
during the layer removal experiment, the frozen-in strains in the plate do not change.
Therefore, we only require the residual temperature eld to numerically simulate the
response of the material to sectioning. Both plates from the previous Section will
be subjected to numerical layer removal.

8.3.1 Original Residual Stress Pro les


The plates are assumed to be long in the longitudinal direction y, and therefore
plane strain conditions apply. To simulate the initial residual stress state in the
plate after cooling, the symmetrical residual temperature pro les (Figure 8.3) are
applied to the same computational mesh as in the previous section, 100  16 CVs.
The plates remain initially \ at" as they develop symmetrical residual stress pro les
(Figure 8.4). The approximate analytical residual stress pro le based on the ln-cos
residual temperature solution is also given and compares well with the nite volume
solution for the fast cooling case.
190
Chapter 8 8.3. Finite Volume Layer Removal Simulations

Figure 8.4: Residual stresses for single and multiple cooling zones cases (analytical
solution from Equation 2.28 with: 1 = 149:398 m 1, T res = 13:317ÆC).
As can be seen in Figure 8.4, the residual stresses are greatly reduced in the slow
cooling regime. The surface compressive stresses are lowered by a factor of 2.45 and
the inner tensile stresses are 2.05 times lower. Such substantial decrease in residual
stresses is accomplished by increasing the total cooling time by only 15%. In the
at plate the bene t of reducing the residual stresses is not obvious. However, if the
residual stresses become asymmetrical during exploitation, the plate would curve.
Symmetrical stress pro les are usually disturbed by a slow creep processes, which are
primarily driven by large residual stresses. In the plate with lower residual stresses
the creep processes are less noticeable. The symmetrical residual stresses in the
plate can be disturbed instantly by sectioning the plate. The resulting deformations
are higher in the plates with higher residual stresses.
The fast and slowly cooled plates are now subjected to layer removals from the top
surface. Each removed layer disturbs the internal force and moment equilibrium.
The net-forces and net-moments contained within the sectioned plates can be de-
191
Chapter 8 8.3. Finite Volume Layer Removal Simulations

Figure 8.5: Net-forces and net-moments contained within the plate as a consequence
of residual stresses.
termined from the residual stress pro les (Figure 8.4) by means of Equations (8.3)
and (8.4). Quantitatively, these functions are presented in Figure 8.5.
The graphs in Figure 8.5 clearly show that the nite volume calculated residual stress
pro les satisfy the equilibrium conditions of zero net-force and zero net-moment,
when no material is removed from the top surface of the plate (z1 = z0 = 6:35 mm).
Since the initial residual stress pro le is symmetrical, the net-force is also zero in
the mid-plane of the plate also. The maximum force and moment is released when
the top quarter of the plate is removed. The rapidly cooled plate exhibits around
2.2 times higher values for the net-moment during layer removal. It is therefore
expected that this plate will curve 2.2 times more after sectioning, when compared
with the slowly cooled plate.

8.3.2 Residual Stresses During Layer Removals


Numerical simulation of layer removal is performed in several stages and by using
several computational meshes. In the rst instance, the top row of cells in the
192
Chapter 8 8.3. Finite Volume Layer Removal Simulations
original 100  16 mesh is removed, thus creating a mesh with 100  15 control
volumes. This reduces the thickness of the plate by 4z = 12:7=16 ' 0:8 mm. The
thickness 4z is refereed to as the layer thickness hereafter. The position of the top
surface z1 is always measured from the original centre of the plate. With one layer
removed, the top surface is at the location z1 = 6:35 0:8 = 5:55 mm. After that
a second layer is removed to create 100  14 control volume mesh. The process of
removing layers is repeated in succession 8 times, when the original thickness of the
plate is halved. In the nal 100  8 mesh, the location of the top surface is described
by z1 = 0.
For every new computational mesh the residual temperature values can be interpo-
lated from the original residual temperature eld (Figure 8.3). If the control volume
sizes are the same for all meshes, the residual temperature values in cell centres are
simply copied from the cells of the original mesh. The transposed residual temper-
ature pro les for sectioned plates are no longer symmetrical, re ecting the frozen-in
strain state in the plate. The residual temperature value is now lower at the top
surface then at the bottom during the layer removal. This case is equivalent to
a cooling case where a plate is subjected to asymmetrical cooling, which is more
intense on the bottom surface. This would result in a \smile" deformation. The
full set of results of the FV layer removal simulations are presented in Figures 8.6
and 8.7. For comparison, the residual stresses are plotted in the same scale for both
plates. Figure 8.8 illustrates the numerically calculated deformed pro les after 2, 4,
6 and 8 layers are removed from the plate surface.
After each removed layer, the maximum stresses in the plate are lowered. Both top
and bottom surfaces remain in compression throughout. The stresses on the bottom
surface are lowered after the layer is removed. However, the compressive stresses
below the removed layer increase. The opposing stress increments on the top and

193
Chapter 8 8.3. Finite Volume Layer Removal Simulations

Figure 8.6: Redistribution of residual stresses xx in the fast cooled plate subjected
to 8 successive layer removals.

Figure 8.7: Redistribution of residual stresses xx in the slowly cooled plate subjected
to 8 successive layer removals.

194
Chapter 8 8.3. Finite Volume Layer Removal Simulations

a)

b)

c)

d)

Figure 8.8: Results of the FV layer removal calculations for the fast and the slow
cooling cases (layer thickness = 0.8 mm): a) 2 layers removed from the top surface,
b) 4 layers removed, c) 6 layers removed, d) 8 layers removed (half of the original
thickness).
195
Chapter 8 8.3. Finite Volume Layer Removal Simulations
the bottom surfaces are the consequence of superimposed bending stresses generated
via the released moment. The stress pro les in the plate do not seem to be highly
asymmetrical during layer removal. However, the detailed quantitative analysis
can prove that asymmetrical components in the stress pro les exist. Asymmetrical
residual stress pro les are accompanied by visible distortions of the plate as shown
in Figure 8.8.
The deformation behaviour of two plates during layer removals is notably di erent.
The fast cooled plate curves more then twice as much as its slowly cooled coun-
terpart. Although the bending moment in both cases reaches a maximum after a
quarter of the thickness is removed, Figure 8.5, the curvature of the plate keeps in-
creasing even further when more layers are removed. Relative to it, Figure 8.8b, the
curvature doubles after the half-thickness is removed, Figure 8.8d. This is caused
by the fact that the bending resistance of the sectioned plate decreases more rapidly
with the removed layers than the calculated bending moment from Figure 8.5.
The described FV numerical algorithm for the simulation of layer removal can be
employed to any calculated residual temperature eld and would yield very detailed
information on the deformation behaviour of the plate. The assumption of constant
residual stresses in all other directions, apart from the thickness direction, is not
required for the numerical analysis. A plate with a non-uniform residual temperature
pro le would produce a deformed shape with variable curvature along the plate
surface.
In the following section, the FV results will be related to theoretical layer removal
expressions derived previously.

196
Chapter 8 8.3. Finite Volume Layer Removal Simulations
8.3.3 Validation of the Finite Volume Layer Removal Model
The nite volume layer removal results are obtained from the known residual tem-
perature eld. As a result of the nite volume analysis performed on di erent com-
putational meshes, the resulting residual stresses and curvatures are determined.
The initial residual stress pro le and the resulting curvatures are linked with the
general theoretical expression (8.12). In order to test FV curvature results, we shall
back-calculate the initial residual stress pro le using Equation (8.13), which applies
for the plane strain layer removal case. The calculated analytical residual stress
pro le should match the initial FV residual stress distribution. The calculations are
performed for both cases described in the previous section. The curvature pro les
are presented in the left graph in Figure 8.9.

Figure 8.9: The layer removal curvature pro les and corresponding residual stresses
for the fast and slowly cooled plates (the analytical curvature t obtained by solving
the reverse problem, see Section 8.1.3).
All deformed shapes calculated using the FV exhibit the \smile" deformations, and
thus the curvatures are all negative. The initial curvature for both plates are zero.
This can be seen in the graph at the points z1 = z0 . The abscissa z1 = 0 represents
197
Chapter 8 8.3. Finite Volume Layer Removal Simulations
the plate of half the original thickness. The symbols in the graph represent the
curvatures determined by the FV method. The points are well represented with
the linear tting curves, obtained using the least square method [80]. The LSQ
curves are used to describe x(z1 ) function, this function is also di erentiated and
integrated as required in the layer removal formula (8.13). The calculated theo-
retical residual stress curves are plotted in the right graph in Figure 8.9. They
correlate very well with the stresses obtained by the FV analysis. However, some
discrepancies are still noted, although the linear LSQ tting curve seems an almost
perfect mathematical t for the FV curvature results. In Figure 8.9 we have only
determined the residual stress pro le in the top half of the plate. This half was
subjected to layer removals and curvature functions are only valid there. However,
if the plate is initially at, the mirror image of the calculated residual stress pro le
can be assumed for the bottom half of the plate.
Now, we will reverse the residual stress problem: our aim is to calculate the an-
alytical curvature functions from the nite volume residual stress or the residual
temperature results, given in Figures 8.4 and 8.3 respectively. This reversed prob-
lem was formulated previously and was described with Equations (8.14) to (8.17).
The solutions of the reversed problems for fast and slowly cooled plates are repre-
sented by dotted lines in Figure 8.9. Generally, the curves follow the same linear
trend as the calculated LSQ curves. By closely examining the analytical curvature
functions, it is noted that they deviate slightly from the linear LSQ curves. In some
regions in the interval [ 0; z0] they are above, and some other regions below the linear
LSQ curves. Therefore, there are variations of the rst derivatives of the analytical
curvature function d x=dz1. At the same time, the rst derivatives are constant for
the linear LSQ t curves. The rst derivative of the curvature functions is responsi-
ble for the discrepancies in analytical residual stress pro les from Figure 8.9. In this
case the discrepancies are minor, but they exist. The problem may become more
198
Chapter 8 8.3. Finite Volume Layer Removal Simulations
signi cant when residual stresses are calculated from the experimentally measured
curvature data, where even small experimental errors can lead to inaccurate residual
stresses.
We have demonstrated here that the nite volume method can be eÆciently used
to simulate the layer removal technique and that the available analytical solutions
are fully recovered. Each layer removal simulation took typically between 1 to 3
minutes of CPU time to calculate the deformed pro les (Intel Pentium 166 MHz).
More importantly, it was demonstrated that there are numerous potential sources of
errors if residual stresses in the plate are measured using the layer removal method.
The bene ts of initially slow and later fast cooling in the plate manufacture have
also been illustrated.
Other experimental methods for measuring residual stresses su er from similar prob-
lems as the layer removal method. The stresses are measured indirectly and the
complex stress conversion algorithms are needed to obtain the residual stresses in
the material. Another popular semi-destructive residual stress measurement method
is the hole drilling method. A comprehensive overview on the hole drilling methods
can be found in [55, 89].

199
Chapter 9
Summary and Conclusions
The objective of this study was to analyse thermally induced residual stresses in
polymeric materials and the e ects they have on the nal polymeric product. Since
most polymeric products are subjected to high temperatures at one stage of the
manufacture, and later cooled to ambient conditions, residual stresses are inevitably
generated in the material. During non-uniform solidi cation the strains are frozen in
the material, resulting in residual stresses after cooling to ambient temperature. The
prediction of frozen-in strains from the thermal history data is the main contribution
of this work. All the principal ideas of the residual stress algorithm are discussed for
at plate geometries, since the industrial plate casting process is the focus of this
research.
Polymeric materials are viscoelastic materials { their response to mechanical and
thermal loads is time dependent. At higher temperatures the time dependence is
more pronounced, and at lower temperatures the response is predominantly elastic.
During thermal processing, the viscoelastic properties of material change continu-
ously with the changing temperature. However, around the glass transition tem-
perature Tg the changes are most pronounced. This is when material solidi es and
the majority of strains become frozen in the material, thus resulting the residual
stresses after processing. The lled PMMA material which is analysed here, has a
200
Chapter 9 Summary and Conclusions
sharp transition region and its viscoelastic behaviour and the generation mechanism
of frozen-in strains can be simpli ed. The instant freezing idea is adopted, which
postulates that the strain is frozen in the material at the solidi cation temperature
Tg in its entirety. Once the solidi cation is complete, the frozen-in strain patterns
are locked in the material, and future residual stress state is determined.
The instant freezing model is mathematically analysed on a one-dimensional in nite
plate geometry subjected to symmetrical thermal loads. As a result of the viscoelas-
tic analysis, the frozen-in strains are found to be proportional to the temperature
gradients at the moment of solidi cation and the coeÆcient of thermal expansion at
the solidi cation temperature Tg . Of all the thermal history only the temperature
gradients at solidi cation feature in the equations describing the frozen-in strain
generation process. Faster cooling, with higher temperature gradients will induce
more strain into the material. The highest solidi cation temperature gradients exist
on cooling surfaces and are zero at points which solidify last. The frozen-in strains
in the sheet can be uni ed into the residual temperature eld by integrating the
temperature gradients at solidi cation across the plate thickness. The material re-
sponds to the residual temperature eld in the same way as to frozen-in strains. This
makes the residual stress problems analogous to thermal stress problems, which are
more commonly solved.
The theoretical ndings from the one-dimensional analysis are transposed to multi-
dimensional solidi cation cases. While the material is solidifying, the solidi cation
front corresponding to isotherm Tg separates the softened and solid material regions.
With the additional heat conducted across the solidi cation front, the front moves to
a new location. The strains induced in the newly solidi ed region are proportional to
solidi cation temperature gradients. The frozen-in strains now need to be integrated
in the direction which is always orthogonal to the solidi cation front to obtain

201
Chapter 9 Summary and Conclusions
the unique solution for the residual temperature eld. The residual stress state is
determined from there.
The residual stress algorithm described above can be solved analytically for only a
handful of one-dimensional cases that are relevant to industry. With the help of
numerical analysis, the residual stresses can be predicted for complex solidi cation
problems, including 3-D geometries. The Finite Volume method was chosen to solve
the continuum mechanics equations governing momentum and energy. With the
proposed frozen-in strain model, the numerical residual stress analysis consist of
three parts:
 Solution of the transient energy equation for determining the thermal history,
and storing the temperature gradients at solidi cation,
 Evaluation of the frozen-in strains by calculating the residual temperature
eld,
 Solution of the steady state momentum equation with the applied residual
temperature eld as an actual temperature distribution.
The frozen-in strain algorithm is conveniently incorporated into the nite volume
framework. The orthogonal lines are determined for every computational point and
the solidi cation temperature gradients are integrated along them. The numerical
stability of the orthogonal line search algorithm is greatly improved by introducing
the weld line.
The proposed computational method is veri ed numerically and experimentally for
at plate geometries. In numerical validations, the nite volume results are tested
against the available analytical solutions for the orthogonal lines, residual tempera-
ture eld, residual stresses and the distortions of the plate. The basic residual stress
principles are illustrated on several cooling examples. A uniformly heated hot plate
202
Chapter 9 Summary and Conclusions
subjected to symmetrical cooling becomes heavily internally strained, but at, after
cooling. If cooling is asymmetrical, the plate curves. Application of the cooling
medium on the top surface results in a \frown" deformation of the plate.
In order to experimentally validate the residual stress model, several cooling tests are
performed and simulated numerically. In the experiments, uniformly preheated lled
PMMA plates are cooled in a controlled environment. After the experiment, the
plates exhibit distortion due to residual stresses. Accurate temperature histories are
needed for the numerical analysis. During the experiments, a number of temperature
signals are recorded to determine reliable heat transfer conditions for the numerical
analysis. After performing the 3-D residual stress analysis, the model predicted
\frown" distortions of the initially at plate, since the cooling was predominantly
from the top surface. This deformed pro le cannot be directly compared with the
experimentally measured pro le after cooling, since in the experiment the plate was
not at at the moment when cooling commenced. The saddle-like initial pro le
of the plate is encountered on many occasions, which can be measured after the
annealing tests. The corrected experimental and numerically predicted deformed
pro les are found to be in a good agreement. The following two recommendations
for future experimental validation cases can be made:
 Good representation of experimental cooling histories in the numerical anal-
ysis (indirect determination of heat transfer conditions from the measured
temperature signals),
 An unconstrained plate will anneal at temperatures higher then Tg , during
which the initially at plate pro le will distort. The annealing pro le needs
to be assumed as the initial pro le for cooling experiments.
Explanations for saddle-like annealed pro les were not found within this research,
but can probably be attributed to conditions at polymeriastion. A closer analysis of
203
Chapter 9 Summary and Conclusions
the polymerisarion processes and its in uence on mechanical behaviour should help
to identify the annealed pro le phenomena.
A shortcoming in the experimental validation process for the cooling experiments, is
the fact that the residual stress elds generated are highly non-uniform, which makes
them unsuitable for experimental measurements. The layer removal method is not
appropriate since it cannot measure very localised stress elds, but gives average
stress distributions across the thickness. The hole drilling method with the small
hole diameter is prone to experimental errors, since the sensitivity of the method is
very low. The use of bigger hole diameters increases the relieved strains and allows
measurements at greater depths, but averages the residual stresses over larger areas.
A few preliminary measurements using hole drilling methods show residual stress
levels close to numerical predictions [89]. The development of reliable experimental
techniques to measure localised, through-thickness residual stresses would be a useful
continuation of this work. In cases where a crystaline phase is present in the material
(semi-crystalline polymers, crystalline llers) some non-destructive methods, like X-
ray di raction or neutron di raction, may be employed to measure residual stresses
[91]. However, numerically predicted deformed pro les from cooling experiments
presented in Chapter 7, are both qualitatively and quantitatively in good agreement
with the experiments. It is expected that the residual stresses responsible for these
distortions would also correlate well.
The residual stress model can assist cast plate manufacturers to analyse closely the
conditions on the real production line. The modi cations to the cooling process
may be introduced to lower the residual stresses. A two cooling zone approach from
Chapter 8 illustrated a large decrease in the residual stresses at the expense of 15%
longer total cooling times, and hence 15% longer cooling zones. The e ect of residual
stresses is evident when stressed geometry is sectioned, such as in layer removal or

204
Chapter 9 Summary and Conclusions
sanding processes. During sectioning, the asymmetrical residual stress components
become larger, and cause a curvature proportional to residual stresses.
The nite volume method was used to carry out the numerical analysis on several
grounds. The method is characterised by simplicity and strong physical representa-
tion of the governing conservation laws. It is widely used for solving highly non-linear
problems in uid ow, combustion and chemical reactions where it is paramount to
conserve the physical properties. The implementation of the nite volume method
for structural problems began in the late 1980s, when problems of multiphysics
become increasingly important. The problem of multiphysics, with the strong inter-
actions between di erent physical phenomena, like uid-solid coupling, can be more
eÆciently implemented if a single numerical method is used. In that view, the plate
casting simulation process can be extended to include ow of the cooling medium,
polymerisarion reactions and the residual stress model in a single numerical frame-
work. The nite volume method can resolve such large computational domains since
it is very memory eÆcient method, because of the segregated solution algorithm.
Finally, let us reiterate the main points of the frozen-in strain model:
 The model is simple and only basic material properties are used,
 It is general and can be applied to arbitrary cooling conditions and geometries,
 It is accurate,
 Although developed for materials with sharp transition regions, it should give
good approximation of residual stresses for materials with gradual changes in
properties,
 It is simple to implement into other numerical procedures.

205
Appendix A
Measurement of Mechanical
Properties
Modulus of elasticity is de ned as the ratio of uniaxial stress and the corresponding
strain at any point within the linear-elastic region. The way to determine the
modulus is by recording a stress-strain behaviour, which is commonly measured
during the uniaxial tensile test. Initially, the Hooke's law is obeyed and the tensile
modulus can be obtained from the initial slope of the stress-strain curve. In the
polymer testing, it is important to bear in mind the time dependence of mechanical
properties, so tests need to be performed at a constant, prede ned displacement
rate.
Tensile tests are conducted according to standard ASTM D 638M-89 [92]. This
method is suitable for testing plastic materials with a thickness of up to 10 mm.
An electro-mechanical INSTRON machine of a constant crosshead speed type was
used for experiments, Figure A.1. It consists of a xed and a movable member,
each carrying a grip assembly for holding the specimen. For brittle materials it is
important to achieve a state of uniaxial loading in order to avoid the possibility of
introducing additional bending loads on the specimen, which can cause premature
fracture. Self-aligning grips ensure that the longitudinal axis of the specimen is
aligned with the direction of the applied load. In addition, the INSTRON can be
206
Appendix A Measurement of Mechanical Properties

Figure A.1: Schematic representation of the experimental setup used for the uniaxial
tensile tests.
tted with an environmental chamber to enable testing at elevated temperatures.
Load and extension are recorded during the test for subsequent determination of
the stress-strain curve. The applied load is measured by a Load Cell, for which
the maximum load was 10 kN. The distance between two points within the gauge
length is recorded by an axial extensometer. The maximal extension of the ex-
tensometer was 2.5 mm. For temperatures up to 100ÆC the gauge length of the
207
Appendix A Measurement of Mechanical Properties
extensometer was xed to 25 mm, thus enabling the measurement of strain up to
10 %. For temperatures above 100ÆC, the gauge length was decreased to 12.5 mm,
measuring strains up to 20%. Before testing, the extensometer was calibrated using
a calibration micrometer.
Test specimen of type M-I was used, as de ned by ASTM D 638M, and it was
machined according to EN ISO 2818:1996 standard [93]. The specimen is a dumb-
bell shaped and can be employed for testing materials having a thickness between
4 and 10 mm. All tests were performed at a constant crosshead velocity of 10
mm/min. The testing temperature range was between 25 to 110ÆC, and four repeat
tests were performed for each temperature. The temperature dependence of the
Young's modulus obtained from the experiments is presented in Figure A.2.
The results show that material gradually softens as the temperature increases, fol-
lowing an almost linear trend. In the temperature interval from 100 to 110ÆC, the
sti ness of the material substantially decreases. The glass transition temperature Tg

Figure A.2: Temperature dependence of Young's modulus for lled PMMA.

208
Appendix A Measurement of Mechanical Properties
is expected to fall within this temperature interval. At high temperatures it is diÆ-
cult to de ne the the initial elastic region since viscoelastic e ects become dominant,
causing rapid stress relaxation. The stress-strain curve measured at 110ÆC is given
in Figure A.3. Here, the material behaviour approaches that of a non-Newtonian
uid, with stress proportional to the applied strain rate.

Figure A.3: Visco-elastic ow of the material at the elevated temperature.


The time dependency of the Young's modulus was examined by a series of stress re-
laxation tests. A dumbbell specimen was preheated uniformly in the environmental
chamber, and then rapidly loaded to reach the prescribed strain level. The strain
was maintained constant for several hours, and the time-decaying load cell signal
was recorded. The relaxation experiment is highly sensitive to temperature vari-
ations, so the temperature needs to be controlled very accurately, as any thermal
expansion/contraction would in uence the measured load signal. After the test, the
measured signal can be normalised by the initial load to obtain non-dimensional
relaxation curves, Figure A.4.
209
Appendix A Measurement of Mechanical Properties

Figure A.4: Isothermal non-dimensional relaxation curves.

210
Appendix A Measurement of Mechanical Properties
Measured relaxation curves illustrate the time and temperature dependency of the
tensile properties. After 1 hour from the start of the test, the relaxation rate drops
very low at all measuring temperatures, and the relaxation modulus become almost
constant. At low temperatures (20{40ÆC), the remaining modulus of the material,
E1 , is still relatively high, and for the analysed lled PMMA it is approx. 80%
of the instantaneous modulus value. At high temperatures the relaxation rates are
many times faster, Figure A.4.
In [94], the second round of testing for the same lled PMMA material has been
performed. The non-dimensional relaxation curves are tted there with the following
hyperbolic function:
E (t) C + A  t[min]n
E (0)
= C + t[min]n ; (A.1)
where t is the relaxation time measured in minutes and E (0) is the instantaneous
modulus. Temperature-dependent curve t coeÆcients A; C and n are given in Table
A.1. At long relaxation times, t >>, Equation (A.1) reduces into:
E1 = A  E (0): (A.2)
Hence, the long term modulus at room temperature (25ÆC) is: E1 = 7:62 GPa.
This value is used as a long term modulus throughout the Thesis.
T [Æ C] A C n E (0) [GPa]
25 0.789 1.845 0.396 9.66
40 0.771 1.826 0.350 8.86
60 0.555 1.625 0.379 7.23
80 0.325 1.270 0.375 5.80
100 0.039 0.585 0.538 1.96
110 - 0.126 0.564 -
Table A.1: Temperature-dependent coeÆcients for the non-dimensional relaxation
modulus hyperbolic t.
211
Appendix B
Measuring the Speci c Heat
Capacity

The speci c heat capacity is de ned as the amount of thermal energy required
to raise the temperature of unit mass by 1 Kelvin. Modern di erential scanning
calorimeters (DSC) are designed to measure the di erential heat ow, required to
maintain a sample of the material and a reference sample at the same temperature
[95]. The temperature variation is pre-determined and usually increases linearly over
the temperature interval at the prede ned rate (Æ C/min), Figure B.1. In DSC sys-
tems, the sample and the reference element are provided with individual heaters. It
is possible to keep the temperature of the sample the same as the reference sample by
continuos, automatic adjustment of the sample heater power. A signal proportional
to the di erence between the heat input to the sample and the reference sample is
recorded as well as the average temperature of both the sample and the reference.
The operational temperature range for modern di erential scanning calorimeters is
typically from -175 to 725 Æ C.
A method of encapsulation of samples is widely used for the DSC measurement.
The sample is placed in an aluminium pan with a domed lid. The contact surface
between a pan and a sample should be maximised. To achieve this, thin disc-shaped
212
Appendix B Measuring the Speci c Heat Capacity

Figure B.1: Typical DSC output signals and predetermined temperature variation.
specimens are commonly used. When the sample material is subjected to a linear
temperature increase, the rate of heat ow into the sample is proportional to its
instantaneous speci c heat capacity. We may write
dH dT
dt
= m  Cp  p ;
dt
(B.1)
where m is the mass of the sample, Cp is the speci c heat capacity and dTp=dt is
the pre-determined rate of the temperature increase. This equation can be used
to calculate Cp directly, but any error in signal read-out or temperature variation
would reduce the accuracy of the result. To minimise the errors, the same procedure
is repeated with a referenced sample (sapphire) for which the speci c heat capacity
is well-established [96]:
dHs
dt
= ms  Cps  dTdtp : (B.2)
By dividing Equations (B.1) and (B.2), and by regarding dH=dt as the signal am-
plitude, it follows:
Cp (sample) =
Amplitude(sample)  m(sapphire) C (sapphire): (B.3)
Amplitude(sapphire) m(sample) p
213
Appendix B Measuring the Speci c Heat Capacity
The signal amplitude represents the rate of heat ow required to rise the temperature
of the material placed into the calorimeter. The sample material is placed into an
aluminium pan and is covered with a thin aluminium lid. The signal amplitude is
then partly due to the heat ux required to heat up the sample material and partly
to the heat ux necessary to heat up the aluminium container. The latter cannot be
ignored. To measure this e ect, a test with an empty pan with the lid is performed
following the same temperature variation. The measured curve is called the base
line. The sample amplitude then represents di erence between the measured curve
and the baseline, Figure B.1.
Three DSC measurements were performed:
 Filled PMMA specimen (weight=20.850 mg),
 Sapphire reference sample (weight=9.120 mg),
 Empty aluminium pan with the lid.
Measured heat ow signals are presented below:

Figure B.2: Measured heat ow signals using the DSC.

214
Appendix B Measuring the Speci c Heat Capacity
Finally, by using the subtracted heat ow signals form Figure B.2 as the \ampli-
tudes" in Equation (B.3), the temperature-dependent speci c heat capacity for the
sample is determined and presented in Figure B.3.

Figure B.3: Temperature dependent speci c heat capacity.


The calculated speci c heat capacity curve is typical for an amorphous polymer.
There is a noticeable change in slope corresponding to the glass transition tem-
perature, Tg , within the temperature interval from 100 to 110ÆC. This is the same
temperature interval where softening of the material has been observed in mechan-
ical tests.

215
References
[1] W. Knappe. Die festigkeit thermoplastischer kunststo e in abhangigkeit von
den verarbeitungsbedingungen. Kunststo e, 51:562{569, 1961.
[2] P. C. Powell. Engineering with bre-polymer laminates. Chapman & Hall,
London, 1994.
[3] O. C. Zienkiewicz. Computational mechanics today. International Journal for
Numerical Methods in Engineering, 34:9{33, 1992.

[4] Abaqus. http://www.abaqus.com.


[5] Ansys. http://www.ansys.com.
[6] Fluent. http://www. uent.com.
[7] Foam. http://www.nabla.co.uk.
[8] C-mold. http://www.mold ow.com.
[9] Physica. http://physica.gre.ac.uk.
[10] C. Bailey, G. A. Taylor, M. Cross, and P. Chow. Discretisation procedures for
multi-physics phenomena. Journal of Computational and Applied Mathematics,
103:3{17, 1999.
[11] A. Ivankovic. Finite volume methodology of dynamic fracture problems. Com-
puter Modelling and Simulation in Engineering, 4:227{235, 1999.

216
References
[12] M. Cross. Modelling of industrial multi-physics processes { a key role for com-
putational mechanics. IMA Journal of Mathematics Applied in Business and
Industry, 7, 1996.

[13] G. E. Mase. Schaum's outline: Theory and problems of continuum mechanics,


chapter 9: Linear Viscoelasticity. McGraw-Hill, 1970.
[14] N. G. McCrum, C. P. Buckley, and C. B. Bucknall. Principles of polymer
engineering. Oxford University Press, 1988.

[15] I. M. Ward and D. W. Hadley. An introduction to the mechanical properties of


solid polymers, chapter 3: Principles of Linear Viscoelasticity. John Wiley &
Sons, 1993.
[16] G. Stephenson. Mathematical methods for science students. ELBS Longman,
second edition, 1973.
[17] S. Matsuoka. Relaxation phenomena in polymers. Hanser Publishers, 1992.
[18] R. J. Young and P. A. Lovell. Introduction to polymers. Chapman & Hall,
London, second edition, 1991.
[19] F. Schwarzl and A. J. Staverman. Time{temperature dependence of linear
viscoelastic behavior. Journal of Applied Physics, 23:838{843, 1952.
[20] L. H. Adams and E. D. Williamson. The annealing of glass. Journal of the
Franklin Institute, 190:597{631 and 835{870, 1920.

[21] E. H. Lee, T. G. Rogers, and T. C. Woo. Residual stresses in a glass plate


cooled symmetrically from both surfaces. Journal of the American Ceramic
Society, 48:480{487, 1965.

217
References
[22] E. H. Lee and T. G. Rogers. Solution of viscoelastic stress analysis problems
using measured creep or relaxation functions. Journal of Applied Mechanics,
30:127{133, 1963.
[23] R. Muki and E. Sternberg. On transient thermal stresses in viscoelastic ma-
terials with temperature dependent properties. Journal of Applied Mechanics,
28:193{207, 1961.
[24] H. S. Carslaw and J. C. Jaeger. Conduction of heat in solids. Oxford University
Press, second edition, 1959.
[25] R. Gardon. Calculation of temperature distributions in glass plates undergoing
heat-treatment. Journal of The American Ceramic Society, 41:200{209, 1958.
[26] V. R. Voller. Advances in numerical heat transfer, chapter 9: An overview of
numerical methods for solving phase change problems. Taylor & Francis, 1996.
[27] P. Chow and M. Cross. An enthalpy control volume unstructured mesh (CV-
UM) algorithm for solidi cation by conduction only. International Journal for
Numerical Methods in Engineering, 35:1849{1870, 1992.

[28] B. D. Aggarwala and E. Saibel. Tempering stresses in an in nite glass plate.


Physics and Chemistry of Glasses, 2:137{140, 1961.

[29] G. M. Bartenev. Investigation of the tempering of glass. Zhurnal Tekhnicheskoi


Fiziki, 19:1423{1433, 1949. in Russian.

[30] O. S. Narayanaswamy and R. Gardon. Calculation of residual stresses in glass.


Journal of The American Ceramic Society, 52:554{558, 1969.

[31] J. G. Williams. On the prediction of residual stresses in polymers. Plastics and


Rubber Processing and Applications, 1:369{377, 1981.

218
References
[32] A. Ivankovic, V. Tropsa, and J. G. Williams. Finite volume modelling of residual
stresses in cast plastic slabs. In T. Ericsson, M. Oden, and A. Andersson,
editors, The fth international conference on residual stresses, volume 1, pages
392{399, June 16-18 1997. ICRS-5, Linkoping, Sweeden.
[33] V. Tropsa, A. Ivankovic, and J. G. Williams. Predicting residual stresses due to
solidi cation in cast plastic plates. Plastics Rubbers and Composities, 29:468{
474, 2000.
[34] Y. Miyano, M. Shimbo, and T. Kunio. Viscoelastic analysis of residual stress
in quenched thermosetting resin beams. Experimental Mechanics, 22:310{316,
1982.
[35] C. E. Maneschy, Y. Miyano, M. Shimbo, and T. C. Woo. Residual stress analysis
of an epoxy plate subjected to rapid cooling on both surfaces. Experimental
Mechanics, 26:306{312, 1986.

[36] ABAQUS Theory Manual, 1994.

[37] I. Skeist. Epoxy resins. Chapman & Hall, London, 1958.


[38] M. Shimbo, T. Shimizu, and Y. Miyano. Generation mechanism of residual
stresses in thermosetting resins introduced during curing (the reciprocation
law of time{temperature{degree of curing). In Proceedings of the 1994 SEM
Spring Conference on Experimental Mechanics, pages 849{854, June 6-8 1994.
Baltimore, USA.
[39] J. E. Martin and D. Adolf. Constitutive equation for cure induced stresses in
viscoelastic material. Macromolecules, 23:5014{5019, 1990.
[40] H. Kawada and K. Ikegami. Viscoelastic properties of resin for IC plastic
packages and residual stress. JSME International Journal, 35:152{158, 1992.
219
References
[41] H. H. Chiang, C. A. Hieber, and K. K. Wang. A uni ed simulation of the lling
and post lling stages in injection moulding. Part 1: Formulation. Polymer
Engineering and Science, 31:116{124, 1991.

[42] H. H. Chiang, K. Himasekhar, N. Santhanamand, and K. K. Wang. Integrated


simulation of uid ow and heat transfer in injection moulding for prediction
of shrinkage and warpage. Journal of Engineering Materials and Technology,
115:37{47, 1993.
[43] H. H. Chiang, C. A. Hieber, and K. K. Wang. A uni ed simulation of the lling
and post lling stages in injection moulding. Part 2: Experimental veri cation.
Polymer Engineering and Science, 31:125{139, 1991.

[44] R. Y. Chang and S. Y. Chiou. A uni ed K-BKZ model for residual stress anal-
ysis of injection moulded three-dimensional thin shapes. Polymer Engineering
and Science, 35:1733{1747, 1995.

[45] J. F. T. Pittman and I. A. Farah. Comprehensive simulation of cooling process


in plastic pipe manufacture. Plastics, Rubber and Composites Processing and
Applications, 25:305{312, 1996.

[46] L. J. Broutman and S. M. Krishnakumar. Impact strength of polymers: The ef-


fect of thermal treatment and residual stress. Polymer Engineering and Science,
16:74{81, 1976.
[47] M. Akay and S. Ozden. In uence of residual stresses on mechanical and thermal
properties of injection moulded polycarbonate. Plastics, Rubber and Composites
Processing and Applications, 25:138{144, 1996.

[48] R. G. Treuting and W. T. Read. A mechanical determination of biaxial residual


stress in sheet materials. Journal of Applied Physics, 22:130{134, 1951.

220
References
[49] L. J. Broutman and P. So. Residual stresses in polymers and their e ect on
mechanical behaviour. Polymer Engineering and Science, 16:785{791, 1976.
[50] M. Akay and S. Ozden. Assesment of thermal stresses in injection moulded
polycarbonate. Plastics, Rubber and Composites Processing and Applications,
25:145{151, 1996.
[51] R. H. Leggatt and S. M. Tavakoli. Measurement of residual stresses in welded
thermoplastics. Plastics, Rubber and Composites Processing and Applications,
26:222{229, 1997.
[52] M. P. I. M. Eijpe and P. C. Powell. Residual stress evaluation in composites
using a modi ed layer removal method. Composite Structures, 37:335{342,
1997.
[53] M. P. I. M. Eijpe and P. C. Powell. Determination of residual shear stresses in
composites by a modi ed layer removal method. Journal of Materials Science,
33:2019{2026, 1998.
[54] ASTM Standard E837-92. Determining residual stresses by the hole drilling
strain gage method. American Society for Testing and Materials, 1992.

[55] J. Lu, editor. Handbook of measurement of residual stresses. Prentice-Hall,


1996. Society for Experimental Mechanics, Inc.
[56] S. P. Timoshenko and J. N. Goodier. Theory of Elasticity. McGraw-Hill London,
3rd edition, 1970.
[57] L. E. Malvern. Introduction to the mechanics of a continuous medium. Prentice-
Hall, 1969.
[58] R. Aris. Vectors, tensors and the basic equations of uid mechanics. Dover
Publications, 1989.
221
References
[59] A. Galovic. Nauka o toplini 2 (Thermodynamics 2). Sveucilisna naklada, Za-
greb, 1991. in Croatian.
[60] I. N. Bronstejn and K. A. Semendjajev. Matematicki prirucnik (Mathematical
handbook). Tehnicka Knjiga, Zagreb, 1975. in Croatian.

[61] H. Benker. Practical use of Mathcad. Springer, 1999.


[62] G. W. C. Kaye and T. H. Laby. Tables of physical and chemical constants.
Longman Scienti c & Technical, 1986.
[63] L. R. G. Treloar. The physics of rubber elasticity. Clarendon Press, 3 edition,
1975.
[64] V. Tropsa, A. Ivankovic, and J. G. Williams. A novel numerical procedure for
predicting residual stresses in polymeric materials. In S. N. Atluri and F. W.
Brust, editors, Advances in computational engineering and sciences, volume 2,
pages 1926{1931, August 21-24 2000. ICES 2000 Conference, Los Angeles,
USA.
[65] V. Tropsa, A. Ivankovic, and J. G. Williams. Predictive model for solidi cation
induced residual stresses in plastic plates. In M. Cross, editor, The 8th annual
conference of the association for computational mechanics in engineering, pages
222{225, April 16-19 2000. ACME 2000.
[66] V. Tropsa, A. Ivankovic, and J. G. Williams. Predicting residual stresses due
to solidi cation in cast plastics. In A. J. Bottiger, R. Delhez, and E. J. Mitte-
meijer, editors, 5th European Conference on Residual Stresses, pages 259{264,
September 28-30 1999. ECRS5.
[67] I. Demirdzic and A. Ivankovic. Finite volume stress analysis. Lecture notes,
Imperial College, Department of Mechanical Engineering, 1999/2000.
222
References
[68] K. Maneeratana. Development of the nite volume method for non-linear struc-
tural applications. PhD thesis, Imperial College of Science Technology and
Medicine, Department of Mechanical Engineering, 2000.
[69] I. Demirdzic and S. Muzaferija. Finite volume method for stress analysis in
complex domains. International Journal for Numerical Methods in Engineering,
37:3751{3766, 1994.
[70] A. Ivankovic, I. Demirdzic, J. G. Williams, and P. S. Leevers. Application of
the nite volume method to the analysis of dynamic fracture problems. Inter-
national Journal of Fracture, 66:357{371, 1994.

[71] Y. D. Fryer, M. Cross C. Bailey, and C. H. Lai. A control volume procedure


for solving elastic stress-strain equations on an unstructured mesh. Applied
Mathematical Modelling, 15:639{645, 1991.

[72] M. A. Wheel. A geometrically versitale nite volume formulation for plane


elastostatic stress analysis. Journal of Strain Analysis, 31:111{116, 1996.
[73] H. Jasak. Error analysis and estimation for the nite volume method with
applications to uid ows.
PhD thesis, Imperial College of Science Technology
and Medicine, Department of Mechanical Engineering, June 1996.
[74] S. Muzaferija. Adaptive nite volume method for ow prediction using unstruc-
tured meshes and multigrid approach. PhD thesis, Imperial College of Science
Technology and Medicine, Department of Mechanical Engineering, 1994.
[75] N. N. Dioh, A. Ivankovic, P. S. Leevers, and J. G. Williams. Stress wave prop-
agation e ects in split hopkinson pressure bar tests. Proceedings of the Royal
Society of London Series A { Mathematical and Physical Sciences, 449:187{204,
1995.

223
References
[76] A. Ivankovic, I. Demirdzic, J. G. Williams, and P. S. Leevers. Finite volume
method and multigrid acceleration in modelling of rapid crack propagation in
full-scale pipe test. Computational Mechanics, 20:46{52, 1997.
[77] J. H. Ferziger and M. Peric. Computational methods for uid dynamics, chapter
5: Solution of linear equation systems. Springer-Verlag, 1996.
[78] E. Kreyszig. Advanced engineering mathematics. John Wiley & Sons, 1976.
[79] I. Demirdzic. Private communication, 1997.
[80] T. Finney. CALCULUS and the analytic geometry. Longman Higher Education,
2000.
[81] I. Demirdzic, S. Muzaferija, and A. Ivankovic. SA-2D nite volume stress anal-
ysis manual. Manual, Imperial College, Department of Mechanical Engineering,
1999.
[82] D. V. Rosato. Plastics processing data handbook. Chapman & Hall, second
edition, 1997.
[83] B. Kapoor, S. K. Gupta, and A. Varma. Parametric sensitivity of chain poly-
merisation reactors exhibiting the trommsdorf e ect. Polymer Engineering and
Science, 29:1246{1258, 1989.

[84] miniPOD Manual and miniPOD Windows Software Manual, 1996. Supreme
mini-POD data acquisition system.
[85] T. J. Quinn. Temperature, chapter 6: Thermocouples. Academic Press, second
edition, 1983.
[86] H. Martin. Heat and mass transfer between impinging gas jets and solid sur-
faces. Advances in Heat Transfer, 13:1{60, 1977.
224
References
[87] LK G-90C Coordinate Measuring Machine (reference manual). LK-DMIS in-
spection software, version 2.0.
[88] H. G. Weller, G. Tabor, H. Jasak, and C. Fureby. A tensorial approach to com-
putational continuum mechanics using object orientated techniques. Computers
in Physics, 12:620{631, 1998.

[89] N. Tassy. Measuring residual stresses in polymeric materials. 3M nal project,


Imperial College, Department of Mechanical Engineering, June 2000.
[90] R. T. Fenner. Engineering elasticity: application of numerical and analytical
techniques, chapter 7: Plates and shells. Ellis Horwood, 1986.

[91] V. Hauk, A. Troost, and D. Ley. Evaluation of (residual) stresses in semicrys-


talline polymers by X-rays. Advances in Polymer Technology, 7:389{396, 1987.
[92] ASTM Standard D638M-89. Standard test method for tensile properties of
plastics (metric). American Society for Testing and Materials, 1989.

[93] European standard EN ISO 2818. Plastics { preparation of test specimens by


machining. CEN European Committee for Standardisation, 1994.

[94] J. Jagust. Mechanical properties of cast polymeric materials and numerical


simulation of the method for determining residual stresses. 3M nal project,
Imperial College, Department of Mechanical Engineering, June 1997.
[95] Di erential Scanning Calorimeter DSC-2. Users handbook.
[96] Ginnings and Furukawa. Journal of the American Chemical Society, 75:522,
1953.
[97] M. Goossens, F. Mittelbach, and A. Samarin. The LATEX companion. Addison
{ Wesley, 1999.

225

You might also like