You are on page 1of 56

On the shrinkage of metals and its

effect in solidification processing

by

Anders Lagerstedt

Casting of Metals
Royal Institute of Technology
SE-100 44 Stockholm, Sweden

Akademisk avhandling som med tillstnd av Kungliga Tekniska Hgskolan i


Stockholm framlgges till offentlig granskning fr avlggande av teknologie
doktorsexamen fredagen den 10 december 2004 kl. 10.15 i Kollegiesalen,
Kungliga Tekniska Hgskolan, Valhallavgen 79, 100 44 Stockholm
ISSN-1104-7127 TRITA-MG-2004:02
On the shrinkage of metals and its effect in
solidification processing

Anders Lagerstedt
Casting of Metals
Royal Institute of Technology, KTH
SE-100 44 Stockholm, Sweden

Abstract
The shrinkage during solidification of aluminium and iron based alloys has been
studied experimentally and theoretically. The determined shrinkage behaviour has
been used in theoretical evaluation of shrinkage related phenomena during
solidification.
Air gap formation was experimentally studied in cylindrical moulds. Aluminium
based alloys were cast in a cast iron mould while iron based alloys were cast in a
water-cooled copper mould. Displacements and temperatures were measured
throughout the solidification process. The modelling work shows that the effect of
vacancy incorporation during the solidification has to be taken into account in
order to accurately describe the shrinkage.
Crack formation was studied during continuous casting of steel. A model for
prediction of crack locations has been developed and extended to consider non-
equilibrium solidification. The model demonstrates that the shrinkage due to
vacancy condensation is an important parameter to regard when predicting crack
formation.
The centreline segregation was studied, where the contributions from thermal and
solidification shrinkage were analysed theoretically and compared with
experimental findings. In order to compare macrosegregation in continuous
casting and ingot casting, ingots cast with the same steel grade was analysed.
However, the macrosegregation due to A-segregation is driven by the density
difference due to segregation. This is also analysed experimentally as well as
theoretically.
This thesis is based on the following supplements:

Supplement 1
J. Kron, A. Lagerstedt and H. Fredriksson
Measurements and modelling of air gap formation in aluminium base alloys
ISRN KTH-MG-INR-04:02SE TRITA-MG 2004:02
Accepted for publication in International Journal of Cast Metals Research

Supplement 2
A. Lagerstedt, J. Kron, F. Yosef and H. Fredriksson
Measurements and modelling of air gap formation in iron base alloys
ISRN KTH-MG-INR-04:08SE TRITA-MG 2004:08

Supplement 3
J. Kron, A. Lagerstedt and H. Fredriksson
Modelling of air gap formation in solidification processing
International conference on solidification science and processing, Bangalore,
India, November 17-20, 2004

Supplement 4
A. Lagerstedt and H. Fredriksson
A model for prediction of cracks in a solidifying shell
ISRN KTH-MG-INR-04:06SE TRITA-MG 2004:06

Supplement 5
A. Lagerstedt, S. Adolfi and H. Fredriksson
Crack formation during continuous casting of tool steel
ISRN KTH-MG-INR-04:07SE TRITA-MG 2004:07

Supplement 6
A. Lagerstedt and H. Fredriksson
Macrosegregation in continuously cast tool steel
ISRN KTH-MG-INR-04:05SE TRITA-MG 2004:05

Supplement 7
A. Lagerstedt, J. Sarnet, S. Adolfi and H. Fredriksson
Macrosegregation in ingot cast tool steel
ISRN KTH-MG-INR-04:09SE TRITA-MG 2004:09
Contents
1. Introduction .............................................................................................. 1
1.1 Background to shrinkage in solidification processes............................ 1
1.2 Aim and content of the thesis................................................................ 2

2. Modelling of solidification ................................................................... 3

3. Air gap formation ................................................................................... 7


3.1 Introduction ........................................................................................... 7
3.2 Experimental work................................................................................ 8
3.3 Mathematical model............................................................................ 11
3.4 Discussion ........................................................................................... 16

4. Crack formation .................................................................................... 17


4.1 Introduction ......................................................................................... 17
4.2 Experimental work.............................................................................. 18
4.3 Mathematical model............................................................................ 19
4.4 Mathematical evaluation of the experimental work ........................... 22
4.5 Discussion ........................................................................................... 25

5. Macrosegregation ................................................................................. 27
5.1 Introduction ......................................................................................... 27
5.2 Continuous casting.............................................................................. 28
5.2.1 Experimental work .................................................................... 28
5.2.2 Mathematical evaluation............................................................ 30
5.3 Ingot casting ........................................................................................ 33
5.3.1 Experimental work .................................................................... 33
5.3.2 Mathematical evaluation............................................................ 35
5.4 Discussion ........................................................................................... 37

6. Concluding remarks ............................................................................. 39

7. Future work ............................................................................................ 41

Acknowledgements ......................................................................................... 43

Bibliography ..................................................................................................... 45
Chapter 1

Introduction

1.1 Background to shrinkage in solidification processes


During solidification of a metal, the density of the material changes due to cooling
of the metal in both liquid and solid state as well as due to the solid to liquid phase
transformation itself. Phase transformations in the solid state during solidification
may also cause a volume change which will affect the solidification process.
The superheated liquid metal cools to the liquidus temperature where
solidification starts. During the cooling, the melt experiences thermal contraction
due to the lowered temperature. The solidification is also connected to a density
change. Depending on the solidification interval of the solidifying metal, the
shrinkage may occur at a specific temperature or distributed over an interval.
During subsequent cooling of the solidified material, the solid metal experiences
thermal contraction. Uneven cooling of the solid induces stresses and strains in the
material. Phase transformations in the solid state may contribute to the shrinkage
of the solid metal. The peritectic reaction in steel is one such important phase
transformation, where the transformation from ferrite to austenite is accompanied
with a shrinkage.
Knowledge of the shrinkage behaviour is an important parameter when modelling
such parts of the solidification process as air gap formation, crack formation and
strain induced macrosegregation. The formation of an air gap occurs due to the
contraction of the solid shell which is withdrawn from the mould. The formation
of an air gap drastically reduces the heat transport across the mould/metal
interface, and thereby affects the solidification process by the decreased heat
removal. Cracking stems from the low strength of the solidified material in
combination with a stress build-up in the solidified shell. The stress distribution in
the solid shell during casting depends to a large extent on the thermal shrinkage of
the solid material. Macrosegregation occurs due to the movement of enriched
liquid, which in some cases are induced by shrinkage, but are also caused by
density differences arising from the microsegregation. One type of
macrosegregation which is caused by the shrinkage is the centreline segregation
seen in continuous casting.

1
The effects of non-equilibrium solidification have been in focus in several
studies.1-5 The reason for the investigated effects of the non-equilibrium
solidification has been attributed the incorporation of excess levels of vacancies
into the solidified material. The increased amount of vacancies will decrease in the
solid as the excess level diffuse towards, and condense in, the grain boundaries.
Amongst other effects, the incorporation of vacancies is thought to affect the
evolution of latent heat.
The combination of discoveries of the importance of vacancy incorporation into
the solid during freezing,1-5 and the previous knowledge that vacancy
condensation affects the density of a metal,6,7 implies that the condensation of
vacancies also may effect shrinkage related phenomena during solidification.
Research regarding the shrinkage behaviour with specific aim on air gap
formation in aluminium and copper has also shown that the condensation of point
defects which are incorporated into the solid material during solidification is an
important factor to consider when investigating the strain in a solidifying body.8-10

1.2 Aim and content of this thesis


This work aims toward an increased understanding of the shrinkage behaviour
during solidification and at applying the results in order to better explain some
shrinkage related phenomena occurring in solidification processes.
The formation of air gap during solidification has been investigated in supplement
1, 2 and 3. Air gap measurements are performed and the shrinkage behaviour
during solidification is modelled and verified by the experimental data.
Crack formation is a phenomenon highly dependent on shrinkage induced stresses.
In supplement 4 a model is developed for the prediction of cracks. The model is
based on an elastic formulation and compares the calculated stresses with the
thermomechanical properties of the material. In supplement 5 the model is
extended to include the vacancy condensation, proved to be an important factor of
the shrinkage in supplement 1 and 2. The extended model is compared with
experimental values of the crack positions.
The shrinkage occurring during the final solidification of a continuously cast slab
is responsible for an increased concentration of carbon and sulphur in the centre of
the slab, also known as centreline segregation. The centreline segregation is
discussed in supplement 6, where an analysis of the contribution of cooling and
solidification shrinkage to the centreline segregation is made.
Macrosegregation also occur in ingot casting, which is treated in supplement 7.
The main reason of the macrosegregation in ingots is density differences in the
liquid state. A contributing factor to the density difference is the increase of
alloying elements in the remaining liquid during solidification and not a shrinkage
of the material itself.

2
Chapter 2

Modelling of solidification
Modelling of a solidification process involves consideration of the heat transports
in and away from the solidifying and cooling material. Heat is transferred by
conduction in the material, which is described by the Fourier heat equation:
T
C P = k 2T + q (1)
t
is the density, C P is the heat capacity and k is the heat conductivity of the
studied material. T is the temperature and t is the time. The source term, q ,
describes the heat released from within the metal. During equilibrium
solidification the source term is expressed:
0 when T > TL
dg T
q = ( H tab ) S when TS < T < TL (2)
dT t
0 when T < TS

The tabulated latent heat, ( H tab ) , is the heat released during equilibrium
solidification, g S is the fraction solid, TL is the liquidus temperature and TS is the
solidus temperature.
Depending on the material studied the fraction of solid is expressed differently.
When regarding steel, the change of the fraction solidified material may be
calculated by the sine function given by Rogberg:11
T TL
1 sin
dg S 2 T TL
= S
(3)
dT 2
(TL TS )1

In aluminium the fraction solid may be described by Scheils equation, assuming
that the diffusion of alloying elements in the solid phase is slow:

3
1

T Tm 1 k f



g S = 1 (4)
T
L Tm

Tm is the liquidus temperature of the pure substance, and k f is the partition


coefficient.
During solidification under non-equilibrium, vacancies are incorporated into the
solid phase, and thereby increase the energy level of the solid. The vacancies
diffuse towards the grain boundaries and condense. The condensation is
accompanied by a heat release as the solid decreases its energy level toward the
equilibrium state. When modelling non-equilibrium solidification, the source term
in the Fourier heat equation has to include the release of heat due to vacancy
condensation:

0 when T > TL
dg T x 1
q = ( H sol ) S + ( H v ) v when TS < T < TL (5)
dT t t M
( H v ) x 1
v
when T < TS
t M
The heat released during the non-equilibrium solidification will also decrease
compared to the equilibrium solidification since the energy level does not reach
the equilibrium state. The equilibrium value of the latent heat is divided in two
parts were one part accounts for the heat of solidification, ( H sol ) , and one term
accounts for the condensation of vacancies formed during solidification:2
( H tab ) = ( H sol ) + (xvinc xveq )( H v ) (6)

xvinc is the concentration of vacancies incorporated into the solid during


solidification and xveq is the equilibrium concentration of vacancies. ( H v ) is the
heat released due to condensation of vacancies. The separate term for the
condensation of vacancies makes it possible to account for condensation occurring
below the solidus temperature.
The equilibrium concentration of vacancies in the solid phase is varying with the
temperature. The temperature dependence is written:
S v H v

k k T
x =e
eq
v
B
e B
(7)
S v denotes the entropy of the vacancy and k B is the Boltzmann constant.
Due to the supersaturation of vacancies in the solid phase after the solidification
the fraction of vacancies will decrease toward the equilibrium value by diffusion
of vacancies to grain boundaries where they condense. This diffusion process
occurring in the solid state may be described by a relation used to describe

4
diffusion during homogenisation processes. The average vacancy concentration is
here expressed as a squared sine wave function:
4 2 D
t
xv = x + (x x )e
eq inc eq
2 2
(8)
v v v

D is the diffusion constant and is the diffusion distance. In the mushy zone the
diffusion distance changes over time, as new material solidifies on top of already
solidified material. In the model used in this work, the calculations of the
concentration are simplified to consist of the average value between newly
solidified material with the maximum vacancy concentration given by the
supersaturation, and the value given by the equation 8, calculated with a
characteristic length related to the fraction solidified material at that temperature.
The numerical solution of the presented equations may be performed in several
different ways. In this work, the heat diffusion equation was solved with a finite
difference method. The model assumes constant material properties within a small
and limited volume, and the thermal influence of the neighbouring volumes, as
well as any internal heat release, is calculated during each time step. The time step
has to be small enough so that events occurring during solidification are treated
properly. Solidification over narrow solidification interval can present difficulties
because of the problems to resolve the very short times where heat is released.

5
6
Chapter 3

Air gap formation

3.1 Introduction
Air gap has a big influence on the heat transfer between casting and mould. There
are many studies on the connection between variations in the heat transfer
coefficient and the formation of an air gap.12-18
The formation of an air gap starts when the solid metal shell becomes strong
enough to withstand the pressure from the melt and thereby departs from the
mould due to the contraction.19 Before the formation of an air gap, heat transfer
occurs mainly through conduction. When an air gap forms, the heat transfer is
decreased and can be described by superposition of the radiation and conduction
terms.20 For aluminium the heat transfer due to radiation is insignificant compared
to conduction and could be neglected,21 while for steel it has to be taken into
account. Due to the small size of the gap, convection in the gap can be neglected.
The general explanation to the formation of the air gap is the thermal contraction
of the solid shell. The thermal contractions arise due to the decreased temperature
at the surface and resulting thermal gradients. There are, however, experimental
work showing that the thermal shrinkage does not alone explain the size of the
measured airgap.8,9
The mathematical treatment of the shrinkage and its contribution to the air gap
formation may include modelling of viscoplastic-elastic stresses and strains,22
although some investigations claim that an elastic model is sufficient to describe
the strain during solidification.23 The use of more advanced models has so far not
been able to give a more accurate prediction of the size of the air gap.24
In recent work a model which includes the shrinkage of vacancy condensation has
been developed and used in the examination of air gap formation in pure
aluminium, eutectic aluminium-silicon and Al-Cu, and copper based alloys.8-10
The model connects the material shrinkage with the formation and condensation
of vacancies as well as the thermal contraction when the alloys are cast in a
cylindrical mould. The aim of the present work is to extend this model in order to
obtain a more thorough description of the solidification process and the shrinkage

7
causing air gap formation when aluminium and iron based alloys are solidifying
over a solidification interval.

3.2 Experimental work


Casting experiments were conducted in top poured cylindrical moulds. The Al-
based alloys were cast in a mould made from low carbon steel with an outer
diameter of 250 mm and a thickness of 50 mm. The Fe-based alloys were cast in a
water-cooled copper mould with an inner diameter of 250 mm and a thickness of
10 mm. The top and the bottom of the moulds were insulated in order to achieve a
heat flow in radial direction mainly. During the casting experiments the
temperature distribution in the castings were recorded. In the aluminium based
experiments the temperature distribution in the mould was also measured, while in
the iron base experiments the mould temperature along with the temperatures of
the cooling water were measured. Simultaneously the movements of the inner
mould wall and the solidifying metal were measured giving the air gap as a
function of time. The heights of both moulds were 100 mm and all measurements
were made at half the height. The experimental set-up used for air gap
measurements of Fe-base alloys is schematically drawn in figure 1.

Figure 1: Experimental set-up for iron base experiments.

8
An example of the measured temperature distribution in ingot and mould is shown
in figure 2 for an aluminium based alloy. The air gap forming in the same
experiment is shown in figure 3. When the metal is poured, the mould is heated
with a consequent expansion. When the cast metal has developed a solid shell
thick enough to withstand the pressure of the liquid metal, the solid shell contracts
and departs from the mould. At the same time, the continued heating of the mould
causes a continued expansion. The air gap is the distance between the mould and
the metal, and will therefore consist of both movements.
The same type of curves for the temperature distribution and air gap formation in
an iron base alloy are shown in figures 4 and 5. The metal solidifies quickly on the
surface of the water-cooled copper mould and an air gap is seen to form
immediately. The movement of the mould is somewhat different in this case. The
mould is rapidly heated during the filling of the mould and an expansion can be
seen in the measurement of the mould displacement. The lowered heat transfer
due to the air gap formation, together with the strong cooling by the water, causes
a decrease in the mould temperature and the mould consequently contracts, thus
reducing the growth rate of the formed air gap.
Since the castings have been insulated in bottom and top and the core gets rapidly
heated to the temperature of the melt, the heat losses are very small in all other
directions than the radial direction.

800

700

600
Temperature [C]

500

400

300

200

100
0 50 100 150 200 250 300 350 400
Time [s]

Figure 2. Measured temperature distribution during solidification of Al-4.5%Cu.

9
-4
x 10
5

Casting
3
Displacement [m]

Mould
-1

-2
0 50 100 150 200 250 300 350 400
Time [s]

Figure 3. Measured displacements of casting and mould during solidification of Al-4.5%Cu.

1600

1500

1400

1300
Temperature [C]

1200

1100

1000

900

800
0 100 200 300 400 500 600 700 800 900 1000
Time [s]

Figure 4. Measured temperature distribution during solidification of Fe-10%Cu.

10
-4
x 10
20

Casting
15
Displacement [m]

10

Mould

-5
0 100 200 300 400 500 600 700 800 900 1000
Time [s]

Figure 5. Measured displacements of casting and mould during solidification of Fe-10%Cu.

3.3 Mathematical model


The solidification process was described by the numerical solution of the Fourier
heat equation discussed in chapter 2. The heat equation is solved for cylindrical
coordinates in the radial direction with a Dirichlet boundary condition. The
surface temperature was extrapolated from the measured temperature distribution.
In the aluminium base experiments the extrapolation was made by fitting a
polynomial to the measured data. In the experiments with iron base alloys, the
strong cooling caused an air gap to form immediately and therefore a polynomial
fit could not describe the surface temperature since the surface temperature starts
to decrease before the thermocouples closest to the mould experience any
temperature decrease. The extrapolation of the surface temperature was therefore
made by parallel movement of the time-temperature curve of the thermocouple
closest to the mould. Figure 6 shows the extrapolated interface temperature
together with the calculated and measured temperature at the thermocouple in the
melt closest to the core for an aluminium base alloy, and figure 7 contain the
corresponding measured, extrapolated and calculated temperatures for an iron
base alloy.

11
800

750
Calculated and measured temperature [C]

700 Measured temperature


r = 25 mm

650

600
Calculated temperature
r = 25 mm
550
Interface temperature

500

450
0 50 100 150 200 250 300 350 400
Time [s]

Figure 6. Measured and calculated cooling curves from experiment with Al-4.5%Cu.

1600

1500
Measured temperature
r = 40 mm
Calculated and measured temperature [C]

1400
Calculated temperature
r = 40 mm
1300

1200
Measured temperature
r = 110 mm
1100

Calculated temperature
1000 r = 109 mm

900
Interface temperature

800

700
0 100 200 300 400 500 600 700 800 900 1000
Time [s]

Figure 7. Measured and calculated cooling curves from experiment with Fe-10%Cu.

12
In the thermomechanical models the shrinkage is described by the superposition of
elastic strain and shrinkage due to vacancy condensation. Relations to calculate
the total shrinkage are given in equations 9 and 10 coming from Hookes law and
a volume balance describing the shrinkage achieved by the ordering of the atoms.
The total displacement of the outer surface of the casting is then the sum of the
thermal displacement, uth , and the vacancy induced displacement, uv .

(1 ) Trdr +
(1 )r 2 + (1 + )a 2 R
uth =
r a R a
2 2
a
Trdr

(9)

2 u v R u v2
xv = (10)
R2 a2
In equation 9, is the thermal elongation coefficient and is Poissons number.
The geometry is described by the outer radius, R , the location of the solidification
front, a, and the spatial parameter, r.
The calculations of the thermal strain were based on the results from the
modelling of the solidification process. In order to model the process of vacancy
condensation and its effect on the displacement of the casting, the fraction of
vacancies incorporated in the solid at the solidification front must be known. In
the case of the experiments with aluminium alloys, the incorporated fractions were
estimated by comparing the measured undercooling with data given by
Fredriksson and Emi.2 During the evaluation of the iron based alloys, the fraction
of incorporated vacancies were assumed to be 10 times the equilibrium
concentration at the solidus temperature. Examples of the resulting calculated
displacements of the solidifying ingots are shown for Al-4.5%Cu in figure 8, and
for a Fe-10%Cu in figure 9.

13
-4
x 10
4

Measured displacement

3
Displacement [m]

Thermal displacement

Total displacement

Vacancy displacement

0
0 50 100 150 200 250 300 350 400
Time [s]

Figure 8. Calculated and measured displacements of the Al-4.5%Cu casting.

-3
x 10
2

1.8 Measured displacement

1.6 Total displacement

1.4 Thermal displacement


Displacement [m]

1.2

0.8

0.6

0.4
Vacancy displacement
0.2

0
0 100 200 300 400 500 600 700 800 900 1000
Time [s]

Figure 9. Measured and calculated displacement of Fe-10%Cu casting.

14
In commercially available softwares for simulation of solidification processes, a
boundary condition which usually has to be given is the heat transfer coefficient.
The heat transfer coefficient may be evaluated from the temperature
measurements in the experiments. The amount of heat transported over the
metal/mould interface is described by equation 11, where both conduction and
radiation is included.
dQ
dt
= Ah(Tmet
i i
(
Tmi ) + A (Tmet )
)4 (Tmi )4 = Aheff (Tmeti Tmi ) (11)

A is the contact area at the interface, h is the conduction heat transfer coefficient,
is the emissivity and is the Stefan-Boltzmann constant. The superscript i
denotes interface and the subscrips met and m indicate metal and mould
respectively. The same amount of heat has to be transported away by the mould.
In the aluminium experiments, this transport is given by the temperature gradient
in the mould, as expressed by equation 12. In the experiments concerning iron
base alloys the heat transported away by the mould is expressed in terms of the
temperature difference and flow of the cooling water and the temperature of the
mould.
dQ dT
= k m (12)
dt dr i

dQ dT dV
= Vm m C Pm m + w w C Pw Tw (13)
dt dt dt
V , k , and C P denote the volume, the heat conductivity, the density and the
heat capacity respectively and the subscript w indicates water. Tw is the
temperature change of the cooling water. The contribution to the heat transfer
from radiation can be considered small compared to the heat transfer by
conduction for the aluminium alloys,21 but not for the iron base alloys. The heat
transfer coefficients are presented in figure 10. In the aluminium case, the heat
transfer coefficient first shows a rapid increase, and after reaching a maximum
value a decrease is seen. The first increase is deduced to come from the
extrapolation of the temperate gradient in the mould, and can be ignored. In the
real case the heat transfer is likely to have the maximum value from the beginning
of the solidification. The decrease of the heat transfer coefficient value coincides
with the measured formation of an air gap. In the case of iron solidification in a
water cooled copper mould, the initial solidification is very rapid and the
maximum value can not be distinguished. This is a result of the extrapolation of
the metal interface temperature as well as the fact that solidification starts during
the filling procedure.

15
2000

1800

1600 Al - 4.5%Cu
Heat transfer coefficient [W/m2K]

1400

1200

1000

800

600
Fe - 10%Cu
400

200

0
0 50 100 150 200 250 300 350 400
Time [s]

Figure 10. Evaluated heat transfer coefficients.

3.4 Discussion
Air gap measurements have been performed on several aluminium and iron based
alloys. The modelling work was based on findings that vacancy condensation
occur during a solidification process,1-5 and that the vacancies also influence the
density.6,7 Air gap measurements performed earlier on pure aluminium, eutectic
aluminium and copper have shown that vacancy condensation influences the air
gap formation.8-10
Although good agreement with the experimental data was obtained, the results
from the modelling work show that the influence of the vacancy condensation is
less pronounced in the solidification of iron base alloys. This is attributed to the
fact that the temperature gradient in the solidifying steel is larger. The reason for
this is the stronger cooling since a water-cooled mould was used. Furthermore the
heat conductivity of steel is much lower than for aluminium.
Altogether, the air gap measurements performed on aluminium and iron base
alloys together with the performed calculations including the effect of vacancy
condensation clearly indicate that a non-equilibrium model better explains the
formation of air gap.

16
Chapter 4

Crack formation

4.1 Introduction
During solidification stresses are induced in a solidifying shell due to uneven
shrinkage of the metal. The level of the induced stress depends on the cooling
shrinkage and the temperature gradient in the shell. The formation of cracks
occurs in the cast material as a result of the combined effect of stress build-up
during the solidification process, and low strength of the newly solidified material.
The formation of half way cracks during continuous casting is one such
phenomenon.
In the continuous casting process, the strand is withdrawn during the
solidification, subjecting the strand surface to different cooling conditions as it
leaves the mould, enters different water cooling regions, etc. The reheating of the
strand surface due to decreased cooling has been shown to influence the stress
distribution in the solidified shell so that tensions arise in the inner parts of the
strand.25-27 External loads, including bulging, roll misalignment, bending and
straightening of the strand are other sources for the increase of stresses in a
solidifying shell. Such phenomena have been investigated in several studies.27-30
In order for a crack to form, the stress or strain in the solidifying material has to
exceed a critical value.25
Many existing crack formation models are based on the assumption that the brittle
cracking only may occur while a liquid film is present.31-34 Experiments
concerning high temperature tensile testing of in-situ solidified steel samples
have shown that the transition from brittle to ductile deformation often is found
well below the solidus temperature, and that the transition temperature is
dependent on the cooling rate.35-37
The studies of air gap formation during solidification discussed in the previous
chapter showed that the incorporation of vacancies and the consequent
condensation is an important factor to consider when investigating strains in a
solidifying material. This fact makes it increasingly interesting to investigate how
the consideration of non-equilibrium solidification will affect the cracking
behaviour of a solidifying shell. In supplement 4, an elastic model to predict the

17
crack formation was therefore developed. The model is extended to include the
effects of non-equilibrium solidification and compared to experimentally obtained
values in supplement 5.

4.2 Experimental work


A continuous casting experiment was carried out in slab caster 2 at SSAB
Oxelsund AB. The outline of the caster is seen in figure 11. The caster is
described in more detail in supplement 5 and 6.

Figure 11. Outline of the casting machine used in plant trials.

During the experiment the casting speed was changed so that two different casting
speeds could be investigated in one heat. However, in order to obtain steady-state
conditions, each of the two casting speeds had to be kept constant during a length
equivalent to the metallurgical length. The experiment was started with a casting
speed of 0.66 m/min and after 30 m of cast strand the withdrawal rate was
increased to 0.78 m/min. In all, 59 m of strand were cast.
A total of nine samples were cut from the strand, evenly spread along the length of
the strand. The samples were etched and chemically analysed. The structures in all
samples were purely columnar, and contained half way cracks. Samples cast with
different casting speeds exhibited different initiation points and end points for the
cracks. The locations of the initiation and ending points of the cracks are plotted
as functions of the strand length in figure 12. A change in crack appearance is
clearly visible in the macrostructures. This change of crack appearance is called

18
inner crack initiation, while the outer starting position is called outer crack
initiation.

140
Upper side
Lower side

120 Crack end


Distance from strand surface [mm]

100

Inner crack initiation


80

60

40
Outer crack initiation

20
0 10 20 30 40 50
Strand length [m]

Figure 12. Crack locations as function of cast length.

Thermomechanical properties for the tool steel cast in this investigation were
determined by high temperature tensile testing of in-situ solidified samples. The
solidification interval of the alloy was determined by thermal analysis.
The continuous casting experiment was, apart from the crack investigation
presented in supplement 5, also used as base for an analysis of the centreline
segregation presented in supplement 6.

4.3 Mathematical model


The base in determining if, and where, crack formation occurs is the comparison
between the calculated stress profile and the strength of the material. The stress in
the solidifying material has to exceed the strength of the material given by the
temperature at the position. At the same time, the material has to be in a
temperature interval where it is brittle. Hence, in order to form cracks two
conditions have to be fulfilled.
Figure 13 illustrates the connection between the measured thermomechanical
properties and the calculated stress profile. The left side of figure 13 represents the
measured properties of the material, while the right side of the figure consists of

19
the calculated stress distribution. If the stress in the solidifying material exceeds
the measured ultimate tensile strength and this occurs at a temperature above the
ductile to brittle transition temperature, a crack will form. This procedure is
illustrated as path A in figure 13. On the other hand, if the stress reaches the
maximum strength of the material below the transition temperature the material is
ductile and will therefore deform plastically, see path B in figure 13.
The temperature distribution in the solidifying shell was calculated with the
Fourier heat equation in one dimension with Cartesian coordinates as discussed in
chapter 2. The computational grid is illustrated in figure 14.

Figure 13. Schematic illustration of the connection between stress profile and cracking behaviour.

20
Figure 14. Schematic illustration of the calculation grid.

Kristiansson studied the stress and strain in a solidifying material.23 He derived an


analytical expression for the elastic strain in a solidifying shell and concluded by
comparing with numerical models of the strain, that the elastic strain was not
much different from the total strain in a solidifying shell. The expression derived
by Kristiansson23 is employed in the present study, and extended to include the
strain induced by the condensation of vacancies. The model assumes no ferrostatic
pressure at the solidification front and no remelting. Furthermore, the model is
only valid as long as the coherent part of the mushy zones does not meet at the
centreline of the strand. The total strain rate is rewritten as the sum of the elastic
strain rate, the thermal strain rate, and the strain rate coming from the
condensation of vacancies:
e th v
= + + (14)
t t t t
Here, e is the elastic strain, th the thermal strain and v represents the strain
induced by the condensation of vacancies. In equation 14, any volume change
occurring below the coherence temperature, such as phase transformations, may
be added as additional terms. The thermally induced tension is expressed:
th T
= (15)
t t
The effect on the strain rate from the condensing vacancies is considered to be
equal to the rate of change of the vacancy concentration:
v xv
= (16)
t t

21
The resulting strain in the solid shell may now be written as:
t
1 x (t ) T xv
S
T xv
e ( x, t ) = + dx dt (17)
t * x
S (t ) 0 t t t t
t * is the time at which the solidification front, xS , passes a certain position. When
considering equilibrium solidification, the rate of change of the fraction of
vacancies becomes zero and the expression becomes equal to the expression
derived by Kristiansson.23
Once the strain in the solidifying shell is obtained it has to be converted to the
stress. Remembering that the formation of cracks occurs when the material is
brittle, Hookes law is used to transfer strains into stresses. Hookes law will only
give valid results within its limits of validity, which is below the yield stress.
However, when the material is brittle, the yield stress becomes equal to the
ultimate tensile stress.

4.4 Mathematical evaluation of the experimental work


The mathematical evaluation involves solving the Fourier heat equation in one
dimension with a Robin boundary condition. The boundary condition includes the
heat transfer coefficient which describes the heat transport away from the strand.
The heat transfer coefficient varies during solidification. During continuous
casting, the heat transfer coefficient varies between the cooling zones. Figure 15
shows the cooling curves for the surface, centreline and one point midway
between the centre and surface, for the two casting speeds used in the performed
experiment. The heat transfer coefficient decrease in steps as the strand descends
through the casting machine, which is seen on the increase of the surface
temperature as the heat transfer changes. In figure 15 both equilibrium and non-
equilibrium calculations are shown, but the calculated temperatures coincide and
are therefore seen as one line for each of the casting speeds.
It has since long been established that the recalescence of the surface due to the
decreased cooling introduces higher stresses in the solidified shell which may
cause cracking.25-27 The change of the stresses is illustrated in figure 16. The
figure shows an equilibrium calculation for Fe-10%Ni 30 seconds after a change
of the heat transfer coefficient. It is seen that the larger the decrease of the heat
transfer coefficient, the larger the increase of the stress. The ultimate tensile stress,
as given by the temperature profile in the shell, is also inserted in the figure, and it
is seen that even a moderate decrease of the heat transfer coefficient is enough to
increase the stress level above the ultimate tensile stress.

22
1500

1400
Centreline

1300
Temperature [C]

1200
Halfway

1100

1000

900
Surface

800 v = 0.66 m/min


v = 0.78 m/min
Solidification interval
700
0 5 10 15 20 25
Distance below meniscus [m]

Figure 15. Cooling curves at surface, centreline and intermediate point for both casting speeds.

100

50

0
Calculated stress [MPa]

Rupture limit
-50

-100
HTC change to 900

-150 No change

HTC change to 700

-200 HTC change to 600

HTC change to 400


-250
HTC change to 200

-300
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045
Distance from cooling surface [m]

Figure 16. Calculated stress distribution 30 seconds after a change of heat transfer in Fe-10%Ni.

23
Non-equilibrium solidification did not affect the temperature distribution much,
but the condensation of vacancies below the coherence temperature will induce
strains into the solid material. Calculations on the experiment described above are
shown in figure 17, where the right side limit coincides with the centreline. The
figure contain both equilibrium and non-equilibrium calculations of the stress.
Inserted into the figure is also the ultimate tensile strength given by the
temperature distribution in the material. The stress profiles show that non-
equilibrium solidification has a stress exceeding the ultimate tensile stress close to
the solidification front.

100

50

-50 Rupture limit


Stress [MPa]

-100

-150
Non-equilibrium solidification

-200
Equilibrium solidification

-250

-300

-350
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
Distance from cooling surface [m]

Figure 17. Calculated stress distribution for equilibrium and non-equilibrium solidification.

By gathering information of the positions where the calculated stress exceeds the
measured ultimate tensile stress and the material at the same time is brittle, a crack
plot can be produced. Such plots are seen for the higher casting speed in figure 18
calculated for both equilibrium and non-equilibrium solidification, where the grey
surface marks cracking. For comparison the locations for the crack initiation and
the widening of the cracks are marked by arrows at the bottom of the figures. The
starting point of the cracks should be the position where the stress first reaches the
ultimate tensile stress, and the crack will thereafter elongate into the material as
long as the stress is above the critical value.

24
0 0
Change of HTC Change of HTC

Change of HTC Change of HTC

5 5

Distance below meniscus [m]


Distance below meniscus [m]

10 10

15 15

Two-phase region

Two-phase region
20 20

25 25


0 0.05 0.1 0 0.05 0.1
Distance from cooling surface [m] Distance from cooling surface [m]

Figure 18. Calculated crack positions for equilibrium (left) and non-equilibrium (right) solidification
and casting speed 0.78 m/min.

4.5 Discussion
A model to investigate the influence of thermal and vacancy induced shrinkage on
the formation of cracks has been developed. The brittle behaviour of the material
enables an elastic model to be used. Experimental thermomechanical data is used
to predict the positions where the stress becomes critical and a crack forms.
Measurements from the continuously cast strand show that the location of the
cracks is moved towards the cooling surface when the casting speed is increased.
The investigation of the cracks also shows that two different crack propagations
occur during the solidification process. However, the two different cracking areas
are not completely separated but may be distinguished by a clear widening of the
cracks.
Fairly good agreement is obtained for the prediction of the first crack formation,
but the lack of a relaxation model prevents a better prediction of the second crack
location.

25
26
Chapter 5

Macrosegregation

5.1 Introduction
Macrosegregation occurs during solidification due to the redistribution of
segregated liquid. The reasons for this movement of segregated liquid may be
found in density differences of the material due to temperature or compositional
differences as well as contraction of the material due to solidification and thermal
shrinkage.
In continuous casting, the centreline segregation is the most severe
macrosegregation. During the final solidification of a strand, segregated liquid
from above is sucked into the dendrite network in order to compensate for the
shrinkage. The reason for the formation of the centreline segregation has been
attributed to the thermal contraction of the already solidified shell,38,39 but also to
the solidification shrinkage occurring during the final solidification.40
By applying thermal41-43 or mechanical42,44 soft reduction of the strand the
formation of the opening in the centre may be prevented. Electromagnetic stirring
at the end of solidification has also been reported to reduce centreline
segregation.45,46
During ingot casting, the most common types of segregation in ingots are the
positive and negative segregation zone and the channel segregates.47,48 The
positive and negative segregation zone refers to the fact that an area with less
segregated elements normally is found at the bottom of an ingot while the
composition exceeds the nominal composition in the top of the ingot. The
negatively segregated zone in the bottom of the ingot is explained by the
sedimentation of equiaxed crystals formed in the bulk liquid,49 while the positively
segregated area at the top of the ingot stems from the flow of segregated liquid
towards the top of the ingot.50,51
A-segregates are channel segregates which forms due to natural convection in the
mushy zone due to density variations of the melt. A liquid jet stream melts the
dendrite network forming a channel where highly segregated liquid flows upwards
due to its lower density. The channel is normally visible as pencil like segregates.
The A-segregation is also the main reason for the positively segregated zone at the

27
top of large ingots as the channels transfer alloying elements towards the ingot
top.
The formation of A-segregates can be prevented by choosing the alloying
elements so that the density of the liquid remains constant over the whole
solidification interval.49,50,52 Maintaining the density constant will therefore also
prevent the formation of the positive segregation at the top of the ingot.

5.2 Continuous casting

5.2.1 Experimental work


The experimental work regarding macrosegregation in continuous casting is based
on the experimental work discussed in chapter 4. Two casting speeds were used.
The lower withdrawal rate was chosen so that final solidification occurred before
the strand reached the area where mechanical soft reduction is performed and the
higher casting velocity was chosen so that the final solidification occurred inside
the area where mechanical soft reduction was applied. Samples taken from the
strand were used to investigate the segregation as a function of the distance from
the cooling surface. The representative segregation pattern for the lower casting
speed is shown in figure 19, and the segregation pattern for a sample with final
solidification inside the soft reduction area is seen in figure 20. The maximum
centreline segregation was measured in all samples cut from the strand and figure
21 shows the result as function of the strand length. The casting speed was
increased after 30 meters of cast strand, and was thereafter kept constant during
the remainder of the experiment. In the figure, approximately the first 10 m were
cast at a constant lower speed, while a transition area where the metallurgical
length successively moves downward into the soft reduction zone was seen
thereafter. After 30 m all samples were cast at a constant higher speed.

28
1.6

1.5

1.4
Segregation ratio [C/C ]
0

1.3

1.2

Sulphur
1.1

0.9 Carbon

0.8
0 50 100 145 190 240 290
Distance from lower surface [mm]

Figure 19. Measured carbon and sulphur distribution in sample with final solidification without soft
reduction.

1.25

1.2

1.15 Sulphur
Segregation ratio [C/C ]
0

1.1

1.05

Carbon
0.95

0.9
0 50 100 145 190 240 290
Distance from lower surface [mm]

Figure 20. Measured carbon and sulphur distribution in sample with final solidification with soft
reduction.

29
1.7

1.6
Sulphur

1.5
/C ]
0
max
Segregation ratio [C

1.4

1.3

1.2

1.1 Carbon

1
0 5 10 15 20 25 30 35 40 45 50
Strand length [m]

Figure 21. Segregation ratio as function of the position in the strand.

5.2.2 Mathematical evaluation


The heat transport was solved in one dimension with the Fourier heat equation.
The resulting cooling curves were shown in figure 15. The calculated temperature
distribution is used to calculate the shrinkage of the strand during solidification of
the central parts.
The contribution from the shrinkage cooling to the void forming in the centre, can
be calculated as the total shrinkage of the strand thickness during the solidification
of the centreline. The size of the increase of the central volume is the sum of the
thermal shrinkage due to the temperature decrease:
tS x

Lcool = ( dT )dxdt (18)


tL 0

t L and t S denote the time when the centreline reaches the liquidus and solidus
temperatures respectively, x is half the thickness of the strand and the
elongation coefficient.
The solidification shrinkage which may contribute to the centreline segregation is
the shrinkage occurring in the central parts only. This shrinkage does not change
the considered volume, but will cause a liquid flow into the volume due to the
shrinkage:

30
L
Lsol = ( dg S )dx (19)
0

is the solidification shrinkage and L denotes the width of the examined central
volume. In these calculations, the length L was considered to be equal to the
width of the centreline segregation given by the chemical analysis, see figure 19.
Figure 22 illustrates the shrinkage formed, and also the effect of the mechanical
soft reduction. The reduction is assumed to directly affect the central void.

0.3
Soft reduction zone
Shrinkage during solidification of centre [mm]

0.25
Cooling shrinkage
(0.66 m/min)

0.2

Solidification shrinkage
0.15 (0.66 m/min)

Cooling shrinkage
(0.78 m/min)
0.1

Solidification shrinkage
(0.78 m/min)
0.05

0
16 18 20 22 24 26 28
Position below meniscus [m]

Figure 22. Calculated shrinkage of the central volume.

The central void created due to the shrinkage is filled with liquid from above. The
liquid is sucked downward through the dendrite network and is therefore enriched.
The thermal shrinkage adds volume to the central void while the solidification
shrinkage does not contribute, see figure 23. The effect on the segregation is
therefore different. The volume flow is directed towards the shrinkage volume.
When the strand is compressed, the volume decreases its size and segregated
liquid is squeezed out of the volume. Figure 24 shows the results from
calculations on sulphur for the two casting speeds.

31
Lsol Lcool
Figure 23. Schematic illustration of the central volume where centreline segregation occurs.

1.7
Soft reduction zone
Segregation due to cooling shrinkage
(0.66 m/min)
1.6
Centreline segregation ratio of sulphur [C/C ]
0

1.5

1.4

1.3

1.2
Segregation due to solidification shrinkage
(0.66 m/min)

1.1

Segregation due to cooling shrinkage


(0.78 m/min)
1
Segregation due to solidification shrinkage
(0.78 m/min)

0.9
16 18 20 22 24 26 28
Position below meniscus [m]

Figure 24. Calculated centreline segregation of sulphur due to thermal and solidification shrinkage.

32
5.3 Ingot casting

5.3.1 Experimental work


The macrosegregation in one octagonal 12 ton ingot and one rectangular 10 ton
ingot has been investigated. The ingots were cast by uphill casting and the target
composition was identical to that used in the continuous casting experiment. The
cooled ingots were cut in pieces in order to enable sulphur printing and drilling for
testing of chemical composition. Sample drilling was performed over the vertical
cross section in order to obtain the segregation pattern of C and S and the
chemical composition for the rectangular ingot is shown in figure 25. In the figure,
the centreline composition is shown to the left and the concentrations along three
vertical lines from surface to centre are shown to the right. The arrows in the
picture mark the locations of the horizontal lines. A sulphur print of the
corresponding surface is shown in figure 26.
In the rectangular ingot, an increase of the segregation ratio is seen towards the
top of the ingot. The horizontal lines shown in figure 25 show that the
composition is rather even towards the surfaces at all levels. At higher levels the
ingot has a clearly unstable segregation ratio towards the centre of the ingots. The
instability of the horizontal lines toward the centre coincides with the appearance
of A-segregates, seen in the sulphur print in figure 26. At the bottom, the ingot
shows a somewhat increasing negative segregation toward the centre.

1200 2.5
2
0
1.5
C/C

1000
1
Distance from bottom of ingot [mm]

0.5
0 100 200 300 400
800
1.5

600 1
C/C

0.5
400 0 100 200 300 400
1.5

200
0

1
C/C

C
S
0 0.5
0.5 1 1.5 2 2.5 0 100 200 300 400
C/C Distance from centreline [mm]
0

Figure 25. Carbon and sulphur composition in rectangular ingot.

33
Figure 26. Sulphur print of vertical cross-section of rectangular ingot.

34
5.3.2 Mathematical evaluation
A theoretical evaluation of the entire segregation process during solidification of
large ingots is quite complex. Several advanced calculations of macrosegregation
involving continuum formulations have been presented in the literature,52-55
showing reasonable results. These codes are, in a general case, able to predict the
formation of A-segregates.52,55 However, when the ingot size is large, the mesh
size makes it difficult to resolve the channel segregation in this type of
simulations. The calculations which are performed here, utilises a simpler
approach to calculate the influence of the A-segregation on the total
macrosegregation in large ingots.
During solidification, the density of the liquid metal changes due to the
segregation of alloying elements. A decrease of the density causes the liquid to
move upward. The effect of temperature and segregation on density was
investigated by Olsson,56 and his result is used in the calculations together with the
lever rule and Scheils equation in order to describe the density change of the
liquid, and the result is shown in figure 27.
It is to be noted that the segregation of the heavy elements, such as Mo, increases
the density while the segregation of lighter elements decrease the density. A
carefully chosen initial composition could therefore equalize the effect from the
density decreasing element with those elements which increase the density in
order to decrease the density change during solidification and thereby decrease the
macrosegregation.

50

-50
Density change [kg/m 3]

-100

-150

-200

-250

-300
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Fraction solid [-]

Figure 27. Calculated change of liquid density during solidification.

35
The compositional changes due to sedimentation are calculated according to
equations 20 and 21.49 The concentration in the sediment zone, Cised , and the
concentration in the remaining liquid, CiL , is calculated assuming that each part of
settled solid is followed by W parts of liquid and a relative height of the sediment
zone V f . The relative height of the sediment zones was established from the
sulphur prints.
C i0 (k i + W )
Cised = (20)
1 + W (1 k i )V f

C i0
C =L
(21)
i
(1 k i )V f
1
1+W
Ci0 is the initial composition and k i is the partition coefficient of element i.
The effect of the A-segregates can be expressed as a mass balance considering the
concentration of an element in the bulk liquid and the concentration of the same
element in the liquid in the A-segregate as expressed in equation 22.49
dC iL
V = v A AA (CiA CiL ) (22)
dt
The volume flow of the segregated liquid in the A-channels is given by the
velocity, v A , and the total cross sectional area of the A-segregates, AA . The
concentration of the bulk liquid, C iL , is initially given by equation 21. By
comparing a chemical analysis of the A-segregate, C iA , with Scheils equation the
fraction solid was determined to be 0.7 when the A-segregates form, which is also
consistent with observations made by Suzuki and Miyamoto.57
The liquid velocity is derived from Darcys law.58 The total cross sectional area of
the A-segregates will vary with time and is found by relating the A-segregate
density to the location of the solidification front. The density of A-segregates is
estimated from the horizontal sulphur prints. Heat conduction simulations were
performed using a commercially available software, NovaFlow & Solid, in
order to obtain the location of the solidification front as a function of time.
Equation 22 may be solved to reveal the segregation resulting from the flow of
segregated liquid in the A-channels. The result is plotted along with the measured
data for the rectangular ingot in figure 28.

36
1600

Carbon (calculated)
1400
Carbon (measured) Sulphur (measured)
1200

1000
Height [mm]

800
Sulphur (calculated)

600

400

200

0
0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4
Segregation ratio [C/C ]
0

Figure 28. Calculated and measured composition along the centreline in the rectangular ingot.

5.4 Discussion
Ingot casting and continuous casting experiments has been performed with a steel
of the same grade. Macrosegregation occurs in both experiments, but the origin of
the segregation is essentially different.
In the continuous casting experiment, centreline segregation is seen. The
centreline segregation is caused by a transport of liquid downwards along the
centreline during the solidification of the centre of the strand. This is an effect of
the solidification and cooling shrinkage, which acts to part the strand into two
halves by forming a crack in the centre. When mechanical soft reduction is
applied the segregation is prevented.
The structure and macrosegregation in the two ingots cast in this study is similar
to the generally reported segregation pattern,47,48 a negatively segregated zone in
the lower parts of the ingots and an increasing composition towards the top. The
primary source of the macrosegregation in the ingot cast material is the transport
of segregated liquid in the A-segregates.

37
38
Chapter 6

Concluding remarks
Modelling of shrinkage during solidification of aluminium and iron based alloys
have been performed along with experimental work showing that the effect of
vacancy condensation has to be included in modelling work of solidification
processes.
Air gap formation was studied in a cylindrical mould. Displacements and
temperatures were recorded during the solidification process. Consequent
theoretical modelling was done where the effect of non-equilibrium solidification
was included. In iron based alloys cast in a water-cooled copper mould the
vacancy condensation proved to be less significant due to the large temperature
gradients, while the condensation of vacancies contributed largely to the shrinkage
in the solidification of aluminium based alloys where the cast iron mould gave
smaller temperature gradients. The agreement between experimentally determined
displacement and theoretically calculated shrinkage was good.
Crack formation during solidification was investigated. A one-dimensional model
to predict the crack locations was developed and applied on a continuously cast
slab, where crack positions were experimentally determined. Calculations were
performed for both equilibrium and non-equilibrium solidification and the results
showed clearly that vacancy condensation has to be included in order to properly
predict crack formation.
The shrinkage of a continuously cast strand also causes macrosegregation to occur
along the centre of the strand, known as centreline segregation. The contribution
to the centreline segregation by the cooling of the solidified material was
compared to the contribution by the solidification shrinkage, and it was
theoretically found that the cooling shrinkage is the primary cause of the
centreline segregation. Finally, the macrosegregation in the continuously cast steel
was compared to the macrosegregation seen in ingot casting which is caused
primarily by density differences in the liquid.

39
40
Chapter 7

Future work
This thesis has dealt with the shrinkage during the solidification process and some
of the resulting effects. Experimental measurements have been compared to
numerical results that clearly show that the condensation of vacancies does
influence the shrinkage and subsequently the solidification phenomena that are
caused by density changes in the solid state.
Detailed information on how the incorporation of condensation of the vacancies is
an area in which further investigation is needed. The amount of excess vacancies
incorporated into the solid is one such question. A more detailed knowledge of the
behaviour of the vacancies in the mushy zone is another.
The models presented to calculate the shrinkage in the shell are both one-
dimensional although with different coordinate systems. In order to be able to
predict the air gap formation or cracking in systems with more complex
geometries the gained knowledge on the shrinkage behaviour has to be extended
to more generalised formulas.
The model for prediction of cracks may be further extended. The model does not
consider relaxation when cracks are predicted which makes the further prediction
difficult after the point where the ultimate tensile strength has been reached.
Although the simplicity of the model depends on the use of purely elastic models,
another step would be to extend the stress model to regard also creep and
plasticity.
When dealing with numerical calculations, the data inserted into the calculation is
important. Input data includes all traditional material properties which have to be
known for the material studied. Considering the extension of the shrinkage theory
to include the condensation of vacancies, new information is needed in order to
more precisely describe the non-equilibrium solidification process.

41
42
Acknowledgements
During my work with this thesis, I have received a lot of help from different
people and organisations, in the shape of guidance, financing, discussions,
material, general support and plain cheering. I would like to express my deepest
gratitude to all of you:

Prof. Hasse Fredriksson for supervising me throughout this work.

The Swedish Energy Agency, the Swedish Steel Producers Association, the
Brinell Centre and SSAB Oxelsund AB for financial support of the different
parts of this work.

SSAB Oxelsund AB and Scana Steel Bjrneborg AB for support with equipment
and personnel during the experiments.

The members of the Swedish Steel Producers Association committees JK24045,


JK24047 and JK24048 for helpful and interesting discussions.

The members of the committee for comparative studies of ingot cast and
continuously cast tool steel. Christer Offerman, Per Hansson, Jan Sarnet and Bjrn
Widell deserve special mentioning.

The former and present staff at the Casting of metals. I am especially indebted to
Anders Eliasson for always having time for advices and discussions about
experimental work.

My dear friends! Without all my friends there would be a lot less of the relaxing
good times. All of them cant be mentioned here, but I would like to point out a
few names: Magnus Bergstrand, Jonas Bergstrand, Mikko stlund and Jessica
Elfsberg.

The Lagerstedt family, especially mom and dad for endless love and support

My extended families: the Helmrich family, the Kron family and the Paral family.

Finally, Jenny Kron for everything.

43
44
Bibliography
1. S. Berg, J. Dahlstrm, H. Fredriksson; ISIJ International, vol. 35, 1995, pp.
876-885
2. H. Fredriksson, T. Emi; Materials transactions JIM, vol. 39, 1998, pp. 292-
301
3. J. Mahmoudi, H. Fredriksson; Materials transactions JIM, vol. 41, 2000, pp.
1575-1582
4. J. Tinoco, P. Delvasto, O. Quintero, H. Fredriksson; International Journal of
Cast Metals Research, vol. 16, 2003, pp. 53-58
5. J. Fjellstedt, H. Fredriksson; Advanced Engineering Materials, vol. 5, 2003,
pp. 24-32
6. R.O. Simmons, R.W. Balluffi; Physical Review, vol. 117, 1960, pp. 52-61
7. K. Wang, R.R. Reeber; Philosophical Magazine A, vol. 80, 2000, pp. 1629-
1643
8. J. Kron, T. Antonsson, H. Fredriksson; International Journal of Cast Metals
Research, vol. 14, 2002, pp. 275-285
9. J. Kron, H. Fredriksson; ASM International, Advances in Aluminum Casting
Technology II (USA), 2002, pp. 147-153
10. J. Kron, H. Fredriksson; ISRN KTH-MG-INR-04:03SE TRITA-MG
2004:03, Royal Institute of Technology, Stockholm, Sweden, 2004
11. B. Rogberg; Scandinavian Journal of Metallurgy, vol. 12, 1983, pp. 13-21
12. L.J.D. Sully; AFS Transactions, vol. 100, 1976, pp. 735-744
13. K. Ho, R.D. Pehlke; Metallurgical Transactions B, vol. 16B, 1985, pp. 585-
594
14. K. Ho, R.D. Pehlke; AFS Transactions, vol. 80, 1983, pp. 689-698
15. J. Isaac, G.P. Reddy, G.K. Sharma; British Foundryman, vol. 78, 1985, pp.
465-468
16. W.D. Griffiths; Metallurgical and Materials Transactions B, vol. 30 B,
1999, pp. 473-482
17. K.N. Prabhu, J. Campbell; International Journal of Cast Metals Research,
vol. 12, 1999, pp. 137-143
18. J.G. Henzel, Jr., J. Keverian; AFS Transactions, vol. 68, 1960, pp. 373-379
19. B.P. Winter, R.D. Pehlke, P.K. Troyan; AFS Transactions, vol. 91, 1983, pp.
81-88
20. K. Ho, R.D. Pehlke; AFS Transactions, vol. 92, 1984, pp. 587-598
21. Y. Nishida, W. Droste, S. Engler; Metallurgical Transactions B, vol. 17B,
1986, pp. 833-844
22. M. Bellet, F. Decultieux, M. Menai, F. Bay, C. Levaillant, J.-L. Chenot, P.
Schmidt, I.L. Svensson; Metallurgical and Materials Transactions B, vol.
27B, 1996, pp. 81-99

45
23. J.-O. Kristiansson; Stress and strain during solidification, Ph.D.-thesis, ISSN
0346-718X, Chalmers university of technology, Gteborg, Sweden, 1983
24. J. Kron, M. Bellet, A. Ludwig, B. Pustal, A. Wendt, H. Fredriksson;
accepted for publication in International Journal of Cast Metals Research,
2004
25. A. Grill, J.K. Brimacombe, F. Weinberg; Ironmaking and steelmaking, vol.
3, 1976, pp. 38-47
26. K. Sorimachi, J.K. Brimacombe; Ironmaking and steelmaking, vol. 4, 1977,
pp. 240-245
27. M. El-Bealy; Scandinavian Journal of Metallurgy, vol. 24, 1995, pp. 63-80
28. B. Barber, A. Perkins; Ironmaking and Steelmaking, vol. 16, 1989, pp. 406-
411
29. B. Barber, B.M. Leckenby, B.A. Lewis; Ironmaking and Steelmaking, vol.
18, 1991, pp. 431-436
30. K. Miyazawa, K. Schwerdtfeger; Ironmaking and Steelmaking, vol. 6, 1979,
pp. 68-74
31. T.W. Clyne, G.J. Davies; Solidification and casting of metals, The Metals
Society, London, 1979, pp. 275-278,
32. M. Rappaz, J.-M. Drezet, M. Gremaud; Metallurgical and Materials
Transactions A, vol. 30A, 1999, pp. 449-455
33. H.N. Han, J.-E. Lee, T.-J. Yeo, Y.M. Won, K. Kim, K.H. Oh, J.-K. Yoon;
ISIJ International, vol. 39, 1999, pp. 445-454
34. J.-E. Lee, T.-J. Yeo, K.H. Oh, J.-K. Yoon, U.-S. Yoon; Metallurgical and
Materials Transactions A, vol. 31A, 2000, pp. 225-237
35. B. Rogberg; Scandinavian Journal of Metallurgy, vol. 12, 1983, pp. 51-66
36. K. Hansson, H. Fredriksson; Advanced Engineering Materials, vol. 5, 2003,
pp. 66-77
37. T. Antonsson, K. Hansson, H. Fredriksson, Proceedings of the 2003
International Symposium on Liquid Metal Processing and Casting, Nancy,
France, September 21-24, 2003.
38. G. Engstrm, H. Fredriksson, B. Rogberg; Scandinavian Journal of
Metallurgy, vol. 12, 1983, pp. 3-12
39. C.-M. Raihle, H. Fredriksson; Metallurgical and Materials Transactions B,
vol. 25B, 1994, pp. 123-133
40. H. Mori, N. Tanaka, N. Sato, M. Hirai; Transactions ISIJ, vol. 12, 1972, pp.
102-111
41. P. Sivesson, C.-M. Raihle, J. Konttinen; Materials Science and Engineering,
vol. A173, 1993, pp. 299-304
42. C.-M. Raihle, P. Sivesson, H. Fredriksson; Modeling of casting, welding and
advanced solidification processes VI, 1993, pp. 577-584
43. P. Sivesson, G. Hlln, B. Widell; Ironmaking and Steelmaking, vol. 35,
1998, pp. 239-246
44. N. Jacobson, C.-M. Raihle, N. Leskinen; Scandinavian Journal of
Metallurgy , vol. 21, 1992, pp. 172-180

46
45. C.-M. Raihle; TRITA-MAC 0522, Royal Institute of Technology,
Stockholm, Sweden, 1993
46. M. Nakada, K. Mori, S. Nishioka, K. Tsutsumi, H. Murakami, Y. Tsuchida;
ISIJ International, vol. 37, 1997, pp. 358-364
47. E. Marburg; Journal of metals, 1953, pp. 157-172
48. M.C. Flemings; Scandinavian Journal of Metallurgy, vol. 5, 1976, pp. 1-15
49. A. Olsson, R. West, H. Fredriksson; Scandinavian Journal of Metallurgy,
vol. 15, 1986, pp. 104-112
50. J. Wanqi, Z. Yaohe; Metallurgical Transactions B, vol. 20B, 1989, pp. 723-
730
51. Z. Radovic, M. Lalovic, M. Tripkovic, B. Jakic; ISIJ International, vol. 39,
1999, pp. 329-334
52. T. Fujii, D.R. Poirier, M.C. Flemings; Metallurgical Transactions, vol. 10B,
1979, pp. 331-339
53. M.J.M. Krane, F.P. Incropera; Metallurgical and Materials Transactions A,
vol. 26A, 1995, pp. 2329-2339
54. M.C. Schneider, C. Beckermann; Metallurgical and Materials Transactions
A, vol. 26A, 1995, pp. 2373-2388
55. J.P. Gu, C. Beckermann; Metallurgical and Materials Transactions A, vol.
30A, 1999, pp. 1357-1366
56. A. Olsson; Scandinavian Journal of Metallurgy, vol. 10, 1981, pp. 263-271
57. K. Suzuki, T. Miyamoto; Transactions ISIJ, vol. 18, 1978, pp. 80-89
58. R. Mehrabian, M. Keane, M.C. Flemings; Metallurgical transactions, vol. 1,
1970, pp. 1209-1220

47
48

You might also like