Вы находитесь на странице: 1из 234

Numerical simulation of unsteady state heat transfer in

horizontal continuous casting with cyclic withdrawal.

Item type text; Dissertation-Reproduction (electronic)

Authors Gupta, Debabrata.

Publisher The University of Arizona.

Rights Copyright © is held by the author. Digital access to this


material is made possible by the University Libraries,
University of Arizona. Further transmission, reproduction
or presentation (such as public display or performance) of
protected items is prohibited except with permission of the
author.

Downloaded 24-Dec-2017 15:06:42

Link to item http://hdl.handle.net/10150/185352


INFORMATION TO USERS

This manuscript has been reproduced from the microfilm master. UMI
films the text directly from the original or copy submitted. Thus, some
thesis and dissertation copies are in typewriter face, while others may
be from any type of computer printer.

The quality of this reproduc,tion is dependent upon the quality of the


copy submitted. Broken or indistinct print, colored or poor quality
illustrations and photographs, print bleedthrough, substandard margins,
and improper alignment can adversely affect reproduction.

In the unlikely event that the author did not send UMI a complete
manuscript and there are missing pages, these will be noted. Also, if
unauthorized copyright material had to be removed, a note will indicate
the deletion.

Oversize materials (e.g., maps, drawings, charts) are reproduced by


sectioning the original, beginning at the upper left-hand corner and
continuing from left to right in equal sections with small overlaps. Each
original is also photographed in one exposure and is included in
reduced form at the back of the book.

Photographs included in the original manuscript have been reproduced


xerographically in this copy. Higher quality 6" x 9" black and white
photographic prints are available for any photographs or illustrations
appearing in this copy for an additional charge. Contact UMI directly
to order.

U·M·I
University Microfilms International
A Bell & Howell Information Company
300 North Zeeb Road, Ann Arbor, M148106-1346 USA
313/761-4700 800/521-0600
--------- ----------.-- ---~--- ..-- ...•..... -

Order Number 9121540

Numerical simulation of unsteady state heat transfer in


horizontal continuous casting with cyclic withdrawal

Gupta, Debabrata, Ph.D.


The' University of Arizona, 1991

Copyright ©1991 by Gupta, Debabrata. All rights reserved.

U·M·I
300 N. Zeeb Rd.
Ann Arbor, MI 48106
1

NUMERICAL SIMULATION OF UNSTEADY STATE HEAT TRANSFER


IN HORIZONTAL CONTINUOUS CASTING WITH CYCLIC WITHDRAWAL

by
Debabrata Gupta

Copyright 0 Debabrata Gupta 1991

A Dissertation Submitted to the Faculty of the


DEPARTMENT OF MATERIALS SCIENCE AND ENGINEERING
In Partial Fulfillment of the Requirements
For the Degree of
DOCTOR OF PHILOSOPHY

In the Graduate College


THE UNIVERSITY OF ARIZONA

199 1
2
THE UNIVERSITY OF ARIZONA
GRADUATE COLLEGE

As members of the Final Examination Committee, we certify that we have read

the dissertation prepared by Debabrata Gupta

entitled Numerical Simulation of Unsk~ady State Heat Transfer

in Horizontal Continuous Casting with Cyclic Withdrawal

and recommend that it be accepted as fulfilling the dissertation requirement

for the Oegr.ee of Ph.D.

Date r' ,

2../,,\ 1'1
4 / 5 /9/
u Dare • •

Date

Date

Final approval and acceptance of this dissertation is contingent upon the


candidate's submission of the final copy of the dissertation to the Graduate
College.

I hereby certify that I have read this dissertation prepared under my


direction and recommend that it be accepted as fulfilling the dissertation
requirement.

ad~
Dissertation Director Date
3

STATEMENT BY AUTHOR

This dissertation has been submitted in partial


fulfillment of requirements for an advanced degree at The
University of Arizona and is deposited in the University
Library to be made available to borrowers under rules of the
Library.
Brief quotations from this dissertation are allowable
without special permission, provided that accurate
acknowledgement is made. Requests for permission ·for extended
quotation from or reproduction of this manuscript in "whole or
in part may be granted by the author.

SIGNED:
4

ACKNOWLEDGEMENTS

I would like to thank Professsor D.R. Poirier, my Dissertaion


Director, for meticulously editing various drafts of this
dissertation. Thanks are also due to Professors L. J. Demer
and D. C. Lynch for serving in my Final Examination Committee.
Dr. Adela Allen, Assoc. Dean of the Graduate College, has
helped enormously regarding various uni versi ty rules and
regulations.
It would not have been possible to complete this dissertation
without the patience and encourgement provided by Dr. Dennis
Olsen, my supervisor at Motorola. Thanks to him I had, for the
first time since my original advisor left the University, an
honef;:t combination of time and finances required to complete
this work.
5

TABLE OF CONTENTS

Page
LIST OF ILLUSTRATIONS • .. ... 8

LIST OF TABLES ....... 15


ABSTRACT . . '. • 17
1. INTRODUCTION • • • 19
1.1 Principles and Practice of continuous Casting • 19
1.1.1 Vertical continuous casting (VCC) • • • • 21
1.1.2 Horizontal continuous Casting (HCC) • 23
1.1.2.1 Cyclic withdrawal in HCC • • • • 26
1.1.2.2 Surface Defects in HCC • • • • • 27
1.2 Mathematical Modelling of continuous casting • • 30
1.2.1 Mathematical Model for HCC • • • • 30
1.3 Numerical Techniques and Implementation • 32
2• LITERATURE SURVEY • • • • • • • • • • • • • • • • • • 33
2.1 Horizontal continuous Casting • • • • • • 33
2.1.1 Mold and Secondary Cooling Design. • 33
2.1.2 Withdrawal System • • • • • • 35
2.1.3 Typical Operating Condition for HCC • • • 40
2.2 Thermophysical Properties • • • • • • 43
2.2.1 Heat Transfer Coefficients • 43
2.2.1.1 Heat Transfer in the
Mold-Metal Gap • • • • • 43
2.2.1.2 Mold-Water Heat Transfer. • 48
2.2.1.3 Secondary cooling Heat
Transfer • • • • • • • 51
2.2.1.4 Air Cooling • • • • 59
2.2.1.5 Heat Transfer in the Melt-Pool. 60
2.3 Mathematical Modelling of Continuous casting . • 62
2.3.1 vertical Direct Chill Casting
of Mg Alloys. • • • • • 62
2.3.2 Vertical Direct Chill Casting
of Al Alloys • • • • 65
2.3.3 VCC of Steel •••••• 67
2.3.4 HCC of Steel • 68
2.4 Numerical Methods • • • ••••••• 70
2.4.1 Mesh Refinement. 70
2.4.2 Composites • . • • • •• 71
6

TABLE OF CONTENTS--Continued

Page

2.4.3 Reentrant Corners • • • • • • • • • • • • 73


2.4.3.1 Series Solution • • • • • • • • 74
2.4.3.2 Modification of the Finite
Difference Equations •
2.4.3.3 Mesh Refinement near
• .. 77
Discontinuities • • 78
2.4.4 Solidification • • • • • • • • • • • • • 79
2.4.4.1 Explicit Stefan Condition • • • 80
2.4.4.2 Lumped Latent Heat. • • 81
3. DEVELOPMENT OF NUMERICAL TECHNIQUES
· · · 83
Mesh Refinement . .
3.1.1 Test Problem · · · · · · 84
3.1 84

3.1. 2 Methods of Mesh ······


· Refinement · · 85
·
3.1. 3 PDE and FDE after Coordinate
Transformation
· · · · · · ·· ·· · ·
3.1.4 Results and Discussion
. 86
88
3.1. 5 Conclusion .
' ~. I ~

3.2 · · ·II· Interface


Heat Transfer across a· Type ··· · ··
·
93
95
3.2.1 Finite Difference Formulations
3.2.2
at a Type II Interface
Results and Discussion
···· ·104
97

3.2.3 Conclusion
···· · .113

4. MATHEMATICAL MODEL
···· .114

4.1 The Energy Equation • • • • • • • • • • • • • • 114


4. 2 Thermoph~lsical Properties and Nonlinearities • .115
4.3 Boundary Conditions • • • • .118
4.4 Governing Equations • • • • • • 121
4.4.1 Temperature. • • • • • .121
4.4.2 Fraction Solid • • • • .124
4.4.3 Interface Coefficients • • • • 127
4.5 Finite Difference Equations • . 128
4.5.1 X-Implicit . • • • • • 129
4.5.2 R-Imp1icit • • • • • • • • .131
4.5.3 Equations for the X-Implicit
Time Step en ~ n + ~) • • • • • • .132
4.5.4 Equations for the R-Implicit
Time Step en + % ~ n + 1) • • • • •• 143
7

TABLE OF CONTENTS--Continued

Page
5. PROGRAM STRUCTURE AND VALIDATION .... .153
5.1 Program Structure •• • • • • • • • • • • .153
5.1.1 Capabilities • • • • • • • • • • • • • • • 153
5.1.2 Organization of Modules and Programs •• 153
5.1.3 Brief Description of Programs.. • .157
5.2 Verification • • • • • • • • • • • • • • • • • • 160
5.2.1 Continuous casting of Mg-Al Alloy • • • • 160
5.2.1.1 Estimation of Mold-Metal Gap
Heat Transfer Coefficient • • • 161
5.2.1.2 Simulation Results
and Analyses • • • • • • • • • 165
5.2.1.3 Numerical Results
and Analyses. • • • .175
5.2.2 VCC of Fe-0.7% C Steel • • • • • • • 177
5.2.2.1 Numerical Difficulties .179
5.2.2.2 Results and Analyses • • • • • • 181
5.2.2.3 Conclusion. • • • • • .184
6. SIMULATION OF HORIZONTAL CONTINUOUS CASTING •• 185
6.1 Choice of computational Meshes for HCC • • .185
6.2 Model for Shell Formation in HCC • • • • • • • • 188
6.3 Input Data and Process Parameters •• 192
6.4 Results • • • • • • • • • • • • • • • • • • 193
6.4.1 Effect of Cycle Rate • • • • • • • • • • 193
6.4.2 Effect of Superheat. •• • • • • • • 204
6.4.3 Effect of Casting Speed. • • • • .206
6.5 Conclusion. • • • • • .209
6.6 Future Work • • • • • .2l.0
APPENDIX A: MESH REFINEMENT BY COORDINATE
TRANSFORMATION • • • • • • • • • 212
APPENDIX B: LIQUID VELOCITIES AND TEMPERATURES
NEAR THE SECONDARY WITNESS MARK
IN HCC • • • • • • • • • • • • • • 217
REFERENCES ....... • • • 222
8

LIST OF ILLUSTRATIONS

Figure Page

1 Mold and spray arrangement in continuous


casting • • • • • • • • • • • • • • • • • 20
2 Schematic of vertical continuous
casting (VCC) • . • • • • • • • • • • • • • • • 22

3 Horizontal continuous casting (HCC) for ferrous


and nickel-base alloys with cyclic withdrawal,
(a) with water sprays for secondary cooling,
(b) with graphite chill plates • • • 24
4 Mold design and withdrawal pattern for (a) VCC
and (b) HCC • . • • • • • • • • • • • • 25
5 Measured shell thickness in partially
solidified casting [26] • • • • . • . . • 28
6 Shell growth and formation of witness marks
in'HCC (a) to (e) shell profile at various
stages of withdrawal, (f) appearance
of casting surface • • • • • • • • • . • • • • . 29
7 withdrawal cycles without and with pushback
(Figs. 7,8 [16]) • • • • • • • • • • • • • 38

8 Temperature distribution in the mold


for HCC of 115 mm square steel billets
(from reference [18]) ••.••••• . •• 41
9 The local heat transfer rate (i.e., flux)
in the mold (from reference (18). The curve
for the conventional slab is for VCC • • . 42

10 Computed gap thickness and heat transfer


coefficient for VCC of 82 rom dia round
billet of 0.7% C steel at 21.16 rom s-1 [35] •• 47

11 Estimated gap heat transfer coefficients


for HCC of steel billets. The value of
hgap given by the broken line on the upper
curve is assumed to be constant as suggested
l?y hgap in the "contact region" shown
1n F1g. 10 • • • • • . • • • • • • • • • . . • • 50
9

LIST OF ILLUSTRATIONS--Continued

Figure Page
12 Calculated local heat transfer coefficient
for water-cooling of HCC mold based on
data of Miyashita et al. [18] • • • • • • • • • 52
13 Heat transfer coefficients in the secondary
cooling region [49]: measurements carried
out on a billet caster • • • • • • • • • • 57
14 Heat transfer coefficient as function
of billet surface temperature during
spray cooling at various spray
intensities [50] • • • • • • • • • • • • • • • • 58
15 Comparison between calculated (---)
and experimental (--) temperature
profiles for VCC of AZ80A alloy;
billet dia. 16 in.; mold length 9.5 in.;
casting speed 2.1 in min"' (from [30]) • • • • • 63
16 Comparison between calculated isotherms
and measured solid-liquid interface
profile for a 0.1524 m diameter A6063
aluminum ingot cast at 3.81 x 10"3 mos"'
by VCC (from [34]) • • • • • • • • • • • • 66
17 Schematics for one-dimensional heat transfer
in two layer composites with (a) Type I
internal interface having infinite heat
transfer coefficient, (b) Type II internal
interface with finite heat transfer
coefficient • • • • • • • • • • • • • • • 72
18 Computed temperature profiles for the prototype
problem for mesh refinement. Profiles plotted
at time intervals of 0.2 starting at t = o.
(a) uniform mesh: at x = 0.9, AX = 0.1;
(b) Nonuniform mesh with coordinate
transformation: at x = 0.896, AX = 0.0505 • • • 89
10
LIST OF ILLUSTRATIONS--Continued

Figure Page
19 Effect of mesh refinement on percent error
in temperature gradient at x = 1 computed
by implicit finite difference met:hod: Ca) as
a function of the exact temperature gradient;
Cb) as a function of the Mesh Refinement
Factor. C.T. refers to coordinate
transformation and S&V refers to the
method of Sundquist and Veronis [53] • • • • • • 91
20 Effect of mesh refinement on percent error
in temperature computed by implicit finite
difference as a function of the Mesh
Refinement Factor. C.T. and S&V are defined
under Fig. 19 •• • • • • • • • • • • • • • 94
21 Schematic for central difference scheme
with imaginary points iC+1 and iO-1 used
to solve unsteady state heat transfer
in a two-layer composite with an internal
Type II interface • • • • • • • • • • • • • • • 98
22 Unsteady state temperature profiles
for the Type II interface prototype
problem with h~ = 0.001 by Ca) backward
and central ana (b) forward difference
methods at the interface. • • • • • • • • .105
23 Unsteady state temperature profiles
for the Type II interface prototype problem
with h~o = 0.1 by (a) backward and central
and (b) forward difference methods
at the interface • • • • • • • • • • • • • • • • 106
24 Unsteady state temperature profiles
for the Type II interface prototype problem
with h~ = 10.0 by (a) backward and central
and (b) forward difference methods
at the interface • • • • • • • • • • • • . • • • 107
25 Heat fluxes at the interiace for the Type II
interface prototype problem using various
finite difference schemes: (a) at time = 1.0;
(b) at time = 2.0 • • • • • • • . • • • • • • • 109
11

LIST OF ILLUSTRATIONS--Continued

Figure Page

26 Effect of heD on the computed interface


temperatures for the Type II interface
prototype problem using various finite
difference schemes • • • • • • • • • • • • • • • 111
27 Effect of At on interface temperatures
for the Type II interface prototype problem
using various F.D. schemes • • • • • • • • • • • 112
28 Geometry and boundary conditions for
the HCC process. Note the reentrant
corners at P, Q and R • • • • • • • . • • • • • 119
29 various types of nodes in the Finite
Difference Mesh for the HCC problem:
(a) X-implicit equations, (b) R-implicit
equations • • • • • • . • • • • • • • • . • • .130
30 Mold design capability of the continuous
casting software CASTSIM-PC • • • • 154

31 Major components of the continuous casting


simulation software CASTSIM-PC • • • • • • • • • 156

32 Calculated mold-metal gap heat transfer


coefficient for Run AB from Adenis et ale
[30]. Extrapolations used in calculations
are shown as broken lines and explained
in the text • . . • • • • • • • • . . • • • • • 163

33 Simulation of steady-state temperature field


in the billet and mold for Run AB of Adenis
et ale [30] by the ADI method. Method 3 was
used to calculate hoop • • . • • • • • . • • • .166
34 Approach to steady state in the simulation
of the experimental casting of Adenis et
ale [30] with At = 2s. Temperature at
(1) billet center, 44 cm from inlet end,
and (2) billet surface, 25 cm from inlet end •• 167
12
LIST OF ILLUSTRATIONS--Continued

Figure Page
35 Time history of (a) computed heat input/
output and (b) position of liquidus and
non-equilibrium solidus at billet centerline
for simulation of the experimental casting
of Adenis et ale [30] • • • • • • • • • • .169
36 Comparison between measured and computed
temperatures on the billet surface. • • .171
37 comparison between measured and computed
"temperatures 5.08 cm under the billet
surface • • • • • • • • • • • • • • • • .172
38 Comparison between the experimental (Curve 1)
and simulated (Curve 2) liquidus isotherms.
Also shown is the simulated solidus isotherm •• 174
39 Effect of At on the calculated temperature
at selected locations in the billet .176
40 Effect of mesh refinement on computed
temperatures on billet surface, Adenis Run AB .178
41 Calculated temperature plot in the billet
and mold for VCC of Fe-0.7% C steel.
Conditions given in Table 8. Upwind
difference for the advective term in PDE
during X-implicit calculat~0ns • • • • • • • • • 182
42 Scenario for shell detachment, growth,
breaking and healing during cyclic
withdrawal for HCC used in the model:
(a) old shell shifted in the x-direction
by transposing temperature values;
temperatures at gridpoints marked by • set
to 8 It; (b) a minimum in shell thickness
deve!ops at grid column marked by arrow;
(c) velocities in solid set for next
timestep. For liquid velocities see Fig. 51. .190
13
LIST OF ILLUSTRATIONS--Continued

Figure Page

43 Computed shell profile for HCC of 114 mm


dia. Fe-0.7% C steel billet using withdrawal
Cycle 1 in Table 8. Profiles shown at
. (a) beginning of pull, (b) pull at maximum
speed, (c) end of pull, (d) end of pause.
At each time the four isotherms shown are for,
from top to bottom, fraction solid = 0.05,
0.5, 0.75 and 0.95. Arrows divide segments
from successive cycles • • • • • • • • • • • • • 196

44 Computed shell profiles for HCC of 114 mm


dia. Fe-0.7% C steel billet using withdrawal
Cycle 2. Profiles shown at Ca) beginning
of pull, (b) end of pull, (c) end of pause • • • 197
45 Computed shell profiles for HCC of 114 mm
dia. billet of Fe-0.7% C steel billet using
withdrawal Cycle 3. Shell profiles (a), (b),
(c) and (d) correspond to same stages of the
withdrawal cycle as in Fig. 43 • • . . .198
46 Computed shell profiles for HCC of 114 mm
dia. billet of Fe-0.7% C steel billet using
withdrawal cycle 4. Shell profiles (a), (b),
(c) and (d) correspond to same stages of the
withdrawal cycle as in Fig. 43 . . • •• .199

47 The temperature in the billet and the mold


at the end of the pull cycle during HCC
of 114 mm dia. billets of Fe-0.7% C steel
using Cycle 1. Note the hot spot at the
inlet end of the mold/metal interface • • • 202
48 Dependence of subsurface shell temperature
on withdrawal cycle rate during HCC
of 114 mm dia. billet of Fe-0.7% C steel • • • • 205
49 Computed shell profiles for HCC of 114 mm
dia. billet of Fe-0.7% C steel using Cycle 3
and a melt inlet temperature of 1515°C. Shell
profiles (a), (b), (c) and (d) correspond
to the same stages of the withdrawal cycle
as in Fig. 43 • • • • • • • • • • • • • • • • • 207
14
LIST OF ILLUSTRATIONS--Continued

Figure Page
50 computed shell profiles for HCC of 114 rom
dia. billet of Fe-0.7% C steel using Cycle 3
but an average velocity of 2.0 cm s·'. Shell
profiles (a), (b), (c) and (d) correspond
to the same stages of the withdrawal cycle
as in Fig. 43 • • • • • • • • • • • • • • • • • 208
51 Proposed velocity pattern in the liquid
adjacent to the break-ring and the shell
during the pull phase of HCC • • • • • • • • • • 218
15

LIST OF TABLES

Table Page
1 HCC withdrawal Cycles [11]
for Structural Steels • • • • • • • • • • • 39
2 Heat Transfer Coefficients
in the Mold-Metal Gap • • • • • • • • • • • 44
3 Heat Transfer Coefficients
for Mold-Cooling water • • • • • • • • • • • • • 49
4 Heat Transfer Coefficients
For Spray Cooling • • • • • • • 53
5 Thermal Conductivity in the Melt Pool • 60
6 Characteristics of Meshes Used to Evaluate
Mesh Refinement Techniques • • • •• • 92
7 Input Data to Simulate Run AB of Adenis et ale 162
8 Input Data to Simulate VCC
of Fe-0.7%C Steel [35] •• 180
9 Simulations of VCC of Fe-0.7%C Steel • • • • • • 183
10 Input Data to Simulate HCC of Fe-0.7%C Steel •• 194
11 withdrawal Cycles for HCC Simulation • • .201
12 Hot spots in the Mold at the End
of the Pull Cycle • • • • • • • .203
13 Subsurface Temperature of Shell Adjacent
to the Break Ring at the End
of the Pause Period • • • • • • • • • • • • • • 203
16

LIST OF SYMBOLS
c = specific heat of the metal, J kg- 1 K- 1
= heat transfer coefficient for radiation, W m- 2 K- 1
= heat transfer coefficient for spray cooling
W m- 2 K- 1

hv = heat transfer coefficient for stable film


evaporation, W m- 2 K- 1
k = thermal conductivity, W me' K- 1
VW = water flux density, m3 m- 2 s-1

p = density, kg m- 3
Os = surface temperature of the metal, K
De = evaporation temperature of the cooling agent, K
Db = initial bulk temperature of the cooling agent, K
17

ABSTRACT

Solidification during horizontal continuous casting of


low carbon steel billets with cyclic withdrawal was simulated
and the wavy profile of the solidifying shell characteristic
of this process was reproduced. Effects of rate of withdrawal
cycle, superheat and casting speed were determined.
In order to carry out this simulation in a personal
computer, efficient numerical techniques had to be developed
for mesh refinement by coordinate transformation, interfaces
with temperature discontinuities and re-entrant corners. A
flexible means of mesh generation involving polynomials was
also developed. From the transient heat transfer model finite
difference equations peculiar to each gridpoint in the
solution field were derived and solved by the Alternating
Direction Implicit (ADI) method. Graphics software were
developed to view the results with 3-D as well as contour
plots.
The heat transfer model was verified with published
resul ts of vertical continuous casting of Mg alloys and steel.
Due to its ability to deal with interfaces, unlike previous
work, the present model could solve temperature at both
casting and mold simultaneously. A model for the shell
growth, rupture and healing at the break-ring of horizontal
continuous casting molds was incorporated into the heat
18

transfer model. An interesting result of this simulation was

the presence of transient hot spots in the hot face of the


mold. Elimination of such hot spots should aid shell strength
and hence the casting rate. A semi-quantitative dependence of
the depth of the primary witness mark on cycle rate was also
established.
19

CHAPTER 1

INTRODUCTION

continuous casting is a modern method for solidifying


liquid metals and alloys into long semi-finished shapes, e.g.
billets and slabs. This method possesses significant advan-
tages in saving labor and energy input, and improving product
quality and metal yield, over the traditional method of batch
ingot casting. Therefore the former has replaced the latter
to a great extent.

1.1 Principles and Practice of continuous casting


In continuous casting (Fig. 1), a melt is continuously
fed into a water-cooled mold, and the solidified product is
continuously extracted. The solidified strand is cut to
manageable lengths and then reduced by hot and cold working to
refine the microstructure and improve its mechanical and other
properties before being shaped into final form.
Design features, such as the length of the mold and other
cooling devices, depend on the type of metal or alloy being
cast. Aluminum [1] and magnesium [2] and their alloys
liberate large amounts of heat during solidification but have
high thermal conductivities. Therefore, these alloys can be
cast either horizontally or vertically from fixed short molds.
........
....
' ......
......... ..........................
............. ' ............
, ... ...........................
, ............. .
I'i';·;"-'·;·"";·:';';-;·;"·;""
............ ..............
,., ,
......................
,
,, ',
,',',' c',,,, "",' t', " .,. , ' , ' , ' , ' , '
, I

.:,;.:.:.:.:):;:.:;:.~.:.:.:.:.;;

....
I'Jj
.......
......
,.,,.,,,,,,..,.,,.
t.,.,."",,..,:t
.,',.,')'
.. ..
,',',: ,': .,' ,',',' .. '
: . : , . )0')." , ••
. . .':

.. .. .......
IQ .,- I).,.
." Jo.
" ' ',.,.:
, 1:.,"'.'- .
., . "" , "'. ,.
,..,"" ".....
, ) ",),,.,,""":
:),,).).""""
'):)
"
~ •••• ' ••• ' . ' . ' .,' ):",."'.' ,. ,. ,. :T_
-:.:.:.:.:.:. ',.,',',' :::::::::::::
~
.... .... ....
::::::::::::
' ..
:::::::::::: .,',','.' :::::::::::::
.......:::::: ,'.'.',',.
,:,~,:,:=
.. . ..
, , .:::::::::::::
:::::::.
...:.,-.
Q,
:::::: •• ' •.•• ,= :::::::
III :::::: " :.:.:.:
::::::- ':',.,:.' :::::::
::l :E (") .:.:.: ,','.:,' .:.:.:.
Q,
(Jl
a oQ.
<D
s:: ---111>1:
Q.
•••••••••
'. '.' ••••
"0
1"'1
III
'<
III , 3
-
.... O::::J
::::J""",,;:::;:
<D 3 _.
0 S» S»
""" ::::J a.
:;-co
,;.:.:.

• ••
• ••
......
:.:.:.
,
,I
• ••
I


~ 0' 0 =:
.........
I •
,
-"0 o (J) I
'

0."0 ::::J ::::J" '

·::::...
::l ,
IQ <D <D ,.
m """ ::::J

o~~ .....
!D ,
~ c: ....
.... <D Q.
_.
,:
,
I •

....::l •••••••
·.......
""" co
:::J
•••••
o 0--" " ,. ::h
o
....~ "Qoo
tnoUJ
s»3=
:E (")
S» 0
.... Q.
ClL----.....
-g--"-
.... ..
.. ..
:
··....'.•' • " '
.... v
,

~ .'< "0 0. <D 5·S» ' ....


...." :~:: ::h
o o(j)~ """co'< Q ...... :;:• "....,v
.•.
s= 0 0 ....
(Jl o[a ••
•,
o 5· o· C"' ,'.
III
(Jl

....
rt
co,<::::J
Q'::
....
......
,
·., , =:"0
,.
",
~

ClUJ
=Q.
m'a:
....
"
oc:
21

In the vertically aligned direct chill (D.C.) casting process


[3], the maj ori ty of the solidification is effected by a
stream of water that impinges on the casting below the mold.
For horizontal casting of aluminum or copper alloys [4], the
copper molds are lined with graphite to inhibit freezeback
into the metal delivery system. The graphite liners also aid
in stripping the casting from the mold.
Compared to aluminum and magnesium alloys, ferrous [5]
and nickel [6] alloys have moderate heats of solidification
but relatively low thermal conductivities. Therefore,
continuous casting of these alloys require much longer molds,
and vigorous water cooling [7]. Moreover, graphite-lined
molds cannot be used, so that elaborate auxiliary lubrication
and other arrangements are needed for stripping the casting
from the mold.

1.1.1 vertical continuous Casting (VCC)


Low-carbon ferrous alloys are cast by the vertical
continuous casting (VCC) process [7]. In the VCC process
(Fig. 2), the initial solidification is in vertical or
inclined molds followed by secondary solidification effected
by cooling sprays arranged in a bank. The vertical arrange-
ment requires massive machinery and associated structures and
buildings, thereby driving up installation costs. In some
variations of the VCC machine, the partially solidified strand
22

Straight Mold •
water cooled and
oscillated
Two phase 20ne or
Metallurgical length = 10m approx.

Bend Radius = 30 m approx.

Fig. 2 Schematic of vertical continuous casting (vee).


23
is bent into the horizontal axis. However, such a solution is
not possible for certain alloys e.g. stainless steels that are
prone to hot cracking.

1.1.2 Horizontal continuous casting (HCC)


Horizontal continuous casting (HCC) of bars and slabs of
low-carbon ferrous alloys is becoming increasingly common. The
process was invented in the sixties to cast remelted ferrous
scrap into bars for forging [8]. Further development was
carried out mostly by various West German [9-17] and Japanese
[18-23] companies who commercialized HCC, including the
installation of three plants in the U.S.A. by 1984 [24-26].
In HCC (Fig. 3), the casting is extracted horizontally,
and none of the above-mentioned restrictions of vec applies.
Moreover, it is claimed that the incidence of porosity in the
casting is reduced so that HCC can 'be made suitable for direct
shaping without subsequent hot working [26].
In order to strip the partially solidified shell from the
water-cooled mold, an oscillatory relative motion between the
mold and the solidified shell is superimposed on, the with-
drawal velocity. In VCC the mold is oscillated (Fig. 4a).
Oscillating the heavier integral tundish (liquid metal
reservoir) and mold in HCC requires more sophisticated
machinery and is still undergoing development [13,14].
Typical HCC facilities, therefore, use a pull-push withdrawal
24

Q Q Q
, " 10 0 01 ,, " , ~
," ~

", ., ,...,
~, ,
, .,
I' ,
, .,I'

o 0 0
Straight Horizontal optional Spray Pinch Rolls
Mold with Boron Electro Cooling for Cyclic
Nitride Break Ring Magnetic Withdrawal
Mold is water cooled Stirring
and stationary

(a)

10 0 0 1 Q
.. ,........ ~~,';\
}(:/~~~~f. '.: :.: :'.: .~~.:. :i~::.::~~)~{!f.?'f!t[f!~~/~:~:~·:!,·:·~i::·:!::·:!,:·~~··"!::~:~.!~:!,:·~::~·"!::·~:::-::::·~~W:~·~

""i~rr~'" '8
Straight Horizontal
Mold with Boron
Nitride Break Ring
Mold is water cooled
optional
Electro
Magnetic
Stirring
Indirect
Secondary
Pinch Rolls
for Cyclic
Cooling by water Withdrawal
cooled Graphite
'"
and stationary Chill Plates

(b)

Fig. 3 Horizontal continuous casting (HCC) for ferrous and


nickel-base alloys with cyclic withdrawal, (a) with
wa~er sprays for secondary cooling, (b) with graphite
chl.ll plates.
25

Vert1cal
Veloc1ty

T1me
Water 0,0 I-----jl-----t--f---t----...
cooled
copper Cast1ng
mold
I
I Mold
.... l..&-
I •
Negat1ve
Str1p
Compressive stress Induced on the freshly solidified
shell when the mold moves downward faster than the
cast billet prevents the Shell from fracture

( a)

Boron n1trlde Horizontal


Velocity

Growth of the shell and


healing of cracks during
the pause and pushbac!<
Triple POint Water cooled 'Wavy'shell per IOds prevent she 11
copper mold fracture

(b)

Fig. 4 Mold design and withdrawal pattern for (a) VCC and
(b) HCC.
26

(Fig. 4b) of the strand, itself, at 80-180 cycles/min.


[9-12,15-26]. Because of these peculiarities of the mold
design and operation, as well as restrictions in capabilities
of the strand withdrawal machinery, there are limitations in
the production rate wi th HCC. Typical casting speeds for
steel billets are about 2.5 to 3.5 cm· s·1 [26], which are less
than the 5.0 cm· s·1 reported for slab casting by VCC [27].
Nevertheless, production rates attainable by HCC are already
attractive for mini-steel plants. Moreover, with HCC hot
rolling involved in producing rolled thin strip can be
eliminated [28].

x.1.2.1 cyclic withdrawal in HCC


The l'rithdrawal motion in HCC consists of repetitive pull-
push-pause stages that are designed to prevent fracture of the
solidifying shell and to heal cracks if they form [11,17].
The typical HCC mold consists of two important parts
(Fig. 4b): the water cooled copper mold itself, and a break-
ring of boron nitride inserted between the mold and the
refractory delivery tube [8]. The break-ring is a thermal
insulator and prevents the solidifying shell from growing
upstream into the refractory delivery tube.
Taylor and Zalner [26] demonstrated that the solidifying
shell at the -junction between· the break-ring and the mold
(called the "triple point" [18-21]) grows normal to both the
27

mold surface and the step of the break-ring. Consequently,


the shell is thinnest at a short distance downstream of the
triple point, as can be seen in Fig. 5. This thin section of
the shell may fracture and leave behind part of the shell
attached to the break-ring. Liquid fills the gap created by
such a fracture, and a partially filled "hot tear" is formed
on the casting surface (Fig. 6). The "cold shut," shown in
Fig. 6, results when the liquid that fills the gap solidifies
against the downstream shell.

1.1.2.2 Surface Defects in HCC


The "cold shuts" and "hot tears" on billets cast by HCC
are often called witness marks. Normally the witness marks do
not present any problems, but if they are too deep below the
surface then subsequent mechanical working causes the tears to
open up or the folds to form oxidized interfaces that pene-
trate into the casting. Such surface defects must be removed
by expensive grinding operations, or at worst the casting is
rejected. This aspect of HCC, therefore, has received
considerable attention [16,21,25].
Several workers [16,19,21,25] investigated the effect of
various cyclic withdrawal patterns on primary witness marks on
billets of several grades of steel. Increasing the superheat
reduces the depth of the marks. Experimental results also
show that, for withdrawals at 100 cycles per min. or less, the
~ .5
en
fa .4
Z
:x::
o .3
:i:
~
:j .2
W
::I:
CI) .1

2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 36

DISTANCE FROM BREAK-RING,IN.

Fig. 5 Measured shell thickness in partially solidified casting [26].

N
0)
29

~~:
--
- --
-----------------
----
---
(0 )

~
------------- -
(b)

-
...
~
----.::
- (c)
\
-------- - - - ---

~
~~
-~~ (d)

(e)
SECONDARY PRIMARY
WITNESS) WITNESS
MARK MARK
[

(f)

Fig. 6 Shell growth and formation of witness marks in HCC.


(a) to (e) shell profile at various stages of
withdrawal, (f) appearance of casting surface.
30
depth of the primary witness marks can increase drastically.
The period when the solidifying metal is in static contact
with the break-ring increases inversely with the cycle rate,
and the thickness solidified depends on the square root of
time [29]. Therefore, the increase in depth of the witness
marks below a critical cycling rate indicates that an analysis
of the cold-shut problem should include the transient behavior
of the mold and break-ring. Such an analysis could be
beneficial in optimizing the operation of a HCC machine and
reduce the expense of full-scale experimentation.

1.2 Mathematical Modelling of continuous Casting


Mathematical modelling and numerical simulation of
continuous casting have been going on since the 1940s. As to
be expected, most modelling [30-36] has been for VCC rather
than HCC. Selcuk [37] and verwijs and Weckmann [38] addressed
the HCC of aluminum alloys with steady withdrawal. Recently
a few workers [39-41] did modelling of HCC of ferrous alloys
with cyclic withdrawal but omitted important procedural
details.

1.2.1 Mathematical Model for HCC


A model of the HCC process should account for its
peculiarities listed below:
(i) cyclic withdrawal, i.e., the transient nature of
the process;
31
(ii) realistic accounting for the movement of the
solidifying shell, as well as the liquid contained
within:
(iii) solution of temperature in the solidifying casting
as well as the mold assembly, with boundary
conditions imposed only at the exposed surfaces,
(iv) composite nature of the mold assembly with
temperature discontinuities at interfaces between
adjacent parts of the mold:
(v) heat transfer at mobile solidification fronts, as
well as at corners such as at the triple-point and
at th~mold exit.
In this dissertation, algorithms to individually address the
above-mentioned points are presented. The algorithms are
combined to accurately solve the two-dimensional temperature
distribution in continuous castings.
In industrial-scale HCC mold assemblies, the temperature
has been measured very close to the triple-point, as an aid in
moni toring the process. Some authors have reported the
measurement of such temperatures and corresponding qualities
of casting surface [11]. At too high temperatures, laps form
on the surface of the casting because of shell-breakage and
resolidification, whereas at temperatures too low very deep
primary witness marks form. Therefore, a practical use of a
thermal model would be to study the effect of various mold
32

designs and operating conditions on the formation of laps and


witness marks.

1.3 ~umerical Techniques and Implementation


In order to numerically solve the partial differential
equations that constitute the mathematical model for heat
transfer in HCC, a finite difference method (FDM) is used in
this dissertation. As shown recently by Shih and Chen [42],
contrary to established belief, FDMs are viable even for the
complicated shapes encountered in HCC. In addition, FDM
requires less computer and programmer resources than do other
numerical methods based on finite elements. For this
dissertation, it was necessary to use a mid-range personal
computer with limited memory and speed; therefore it was
important to explore and adapt efficient and stable mesh-
refinement techniques associated with the discretization of
the equations. Consequently, efficient computer programs that
may be run on widely available personal computers, were
developed. The programming was done in PASCAL, in order to
take advantage of structured programming and on-screen
graphics available with this programming language.
33

CHAPTER 2

LITERATURE SURVEY

In this chapter the hardware and operation, as well as


options for mathematical modelling and numerical solution of
the thermal field for HCC are examined.

2.1 Horizontal Continuous Casting


Many authors [11,14,16,21,25] have discussed the develop-
ment and operational aspects of HCC, especially for stainless
and alloy steels. The information provided by them is
reviewed under the categories of mold and secondary cooling
design and withdrawal system, with emphasis on operating
parameters and product quality.

2.1.1 Mold and Secondary Cooling Design


Water-cooled moids made of various copper-base alloys
(e.g., Cu-Co-Be and Cu-Ni-Be alloys [26]) with high hardness
(about ~ 95) have a wear resistance superior to Cu-Cr-Zr
alloys for casting specialty steel alloys with no mold
lubrication. These molds are 40 to 80 cm long and are used to
cast rounds and squares of 60 to 200 mm. For casting hot-
strength alloys, the metallostatic head created by the liquid
metal in the tundish is inadequate to ensure contact between
34

the solidified shell and the mold wall. Therefore, in order


to maximize contact and improve heat transfer with the
solidified shell, which shrinks on cooling, molds are tapered
inward at the exit end [17]. On the other hand, as tapering
increases the withdrawal force, an excessive taper can cause
rupture of the shell in alloys with low hot-strength or wide
freezing range.
The break-ring [8,11] is located at the entry end of the
mold and is a major component of mold-assemblies for both
ferrous [11] and nonferrous alloys [37]. The break-ring is
made of a refractory material of intermediate conductivity and
has an internal diameter slightly smaller (by 4 to 30 rom) than
that of the mold. The step thus created forms a discontinuity
in the liquid metal path and "breaks" the upstream growth of
the solidifying shell on to the delivery tube. The initial
formation of the shell at a predictable location (the triple-
point) also makes it feasible to monitor the shell-growth
process, by measuring temperature in this area. Because the
temperature at the triple-point area is so critical, thermal
modelling of HCC must account for the additional complexity
arising from the break-ring.
The break-ring must not react with the liquid metal and
should not adhere to the solidified shell, for good stripping
during withdrawal. Various grades of boron nitride [11] and
35

alloys of Si-AI-O-N with 10 to 20% BN [21] are suitable for


casting low and high alloy steels and stainless grades.
There are at least two methods of providing secondary
cooling for completing the solidification of the cast strand
in HCC processes. TechnicaGuss [9,10,24,26] applies secondary
cooling with graphite lined water-cooled copper blocks. In a
refinement of this technique [16,17], spring-loaded cooling
sections are used to improve contact with the casting and
increase the heat transfer rate. The length of such a cooling
section is about 1.5 m [26]. Other developers of the original
HCC concept use water-sprays for secondary-cooling, as in the
HORICAST process [19].
Electromagnetic stirring (EMS) is applied to solidifying
strands in order to reduce columnar dendrite formation and
thereby avoid center porosity and center cracks. When linear
stirring is used [26], the dendrites broken by the stirring
make equiaxial grains that settle to the bottom of the melt in
the liquid sump, while columnar grains continue to grow in the
top half of the strand. To make a more uniform structure
rotary EMS is used [13].

2.1.2 withdrawal System


The pUll-pause-push-pause cycle during withdrawal is
accomplished with precisely driven pinch-rolls located about
15 m downstream of the secondary cooling. In the pull stroke
36
the strand is accelerated from rest to a steady forward speed
reaching up to 10 m·s·' [16], and fresh liquid flows into the
mold (Fig. 6a). Solidification proceeds from the triple-
point, with the thinnest part of the shell about 1 rom thick
and located at a position intermediate between the break-ring
and the shell solidified in the previous cycle (Fig. 6b and
6c). After a pull of 10 to 20 rom, there is a pause and the
stationary strand solidifies for 0.1 to 0.2 s, so that the
thinnest part can gain enough strength to withstand the
withdrawal stresses (Fig. 6e).
During the pause the solidifying strand undergoes cooling
and contraction. The resulting stress may induce fracture at
the thinnest part of the shell. Apart from its thinness the
load-bearing capacity of the shell at this point is further
compromised by the directionality of the dendrites that are
inclined at an angle to dendrites downstream of this point.
The plane where the two sets of dendrites meet is crack-prone.
-
In order to prevent the shell from cracking at this point, the
strand is pushed back by only 0.38 mm [16] so that incipient
cracks are healed. Another pause period may follow to further
consolidate the shell. An inadequate pause period or push may
result in a fractured shell and breakout of the liquid during
the subsequent pull. Voss-Spilker and Reichelt [11] showed
that for strands of diameter up to 150 rom the push stroke
37
actually allows a 100% increase in average casting speed and
production rate.

Withdrawal cycles have been given by Haissig [16] and


Voss-Spilker and Reichelt [11]. Figure 7 shows velocities for
four different cycles from the work of Haissig [16], and in
Table 1 three different cycles used by Voss-Spilker and
Reichelt [11] are given. In addition to the individual
components of the withdrawal cycle, the cycle rate, itself, is
important and, as mentioned earlier, has received wide

attention [16,21,25]. Depth of the primary witness marks

(cold shuts) vary from 1 to 4 mm, depending on the cycle rate.


For carbon steels, the cracks associated with the cold shuts
disappeared at withdrawal cycle rates of 150 cycles per min.
or more, but for stainless steels operation at 200 cycles per
min. was required [16,21,25] to eliminate the cracks.

Increasing superheat also reduces the witness mark depth.

When casting a 96 mm dia. billet of AISI 310 [16], increasing


the superheat from 10 to 70·C caused a decrease from 5.5 mm to
3.5 mm in the depth of the primary witness marks.
As mentioned before, a thermocouple inserted near the
triple-point in an HCC mold can be very useful in monitoring
the process. Voss-Spilker and Reichelt [11] give an example
of detecting shell rupture. When there is rupture without

complete healing, the strand becomes stationary and grows


38

16

14

12
c

1
.e..
10

8
...
~
0
6
~ End of cycle
4

2
/' t\~,
ri
0.1 0.6 0.7
TIme, •

Stroh Range, Cyc," Velocity (max), Acc.lenJtion (max), PouJ,

1
".
40
Imin
93
metrft/mln
6.23
mot,..hl.
0.75
rm
238
2 SS 113 7.59 1.11 198

..
3 6S
90
127
160
8.49
10.75
1.38
2.22
17S
138

"'Include. pulhbOclc time


5t,ob length I9.Imm Puahboclc 0.38 mm

Fig. 7 withdrawal cycles without and with pushback


(Fig. 8 [16]).
39
Table 1
HCC withdrawal Cycles [11] for structural steels

Sequence withdrawal Cycles


in Movement, Time, Speed, per
CYcle mm s m·min- 1 min.

Case I Pull 15.0 0.25* 3.6


Push 1.0 0.05 1.2
Pause 0 0.2 0
Total 14.0 0.5 1.7 120

Case II Pull 15.0 0.35* 2.6


Push 1.0 0.05 1.2
Pause 0 0.1 0
Total 14.0 0.50 1.7 120

Case III Pull 10.0 0.17* 3.6


,
Push 0.05 0.03 1.2
Pause 0 0.13 0
Total 9.4 0.33 1.7 180

*Assumes infinitely fast acceleration with no ramp time.

beyond the normal thickness at the triple-point. Correspond-


ing to this event, there is a drastic decrease in temperature
(e.g., 30°C in 4 s). ,The break-off of the casting shell may
be repaired by a pause of 10 s when the shell reforms. When
normal withdrawal is resumed after this period, the tempera-
ture close to the triple-point rises, indicating that shell
break-off has been healed.
40
2.1.3 Typical operating Condition for HCC
Though somewhat incomplete, Miyashita et al. [18] have
provided important information regarding heat transfer in HCC
molds. For casting 115 mm sg. carbon steel billets at an
average speed of 2.1 m· min- 1 , they used a mold of high-strength
copper alloy that was 12 mm thick and 600 mm long. No
dimension of the boron nitride break-ring was given by the
authors but from Fig. 13 of Holleis et al. [40] it was
estimated to be 12 mm high and 13 mm long. The mold was
cooled with water, flowing at rates of 2100 to 2500
liter·min. -1 or velocities ranging from about 210 to
250 m·min. -1. The secondary cooling zone was divided into six
sections with a total length of 6 m, and in this zone solid-
ification was completed by water sprays with water flow rates
ranging from 0.3 to 1.0 liter.kg- 1 , depending upon the steel
grade. For the casting rate employed, the flow rate of water
converts to an average flux of 818 to 2728 liter'm- 2 'min. -1 of
water over the surface of the billet under the sprays.
Measured temperatures in the mold (mold thickness is 12 mm)
for casting steel poured at 1550°C under these conditions is
shown in Fig. 8, and the heat flux along the mold length in
Fig. 9. Temperatures in the billet itself, were not reported.
41

500r-----------------------------------

I
t= 12 mm
Mold l =600 mm
_ 400
~. Material: D-Cup

-
oU

a>
Costing velocity: 2.1 m/min.

Mold cooling {o 21 00 lim in.

-5
o
L-
a>
a.
300
\
.,\ °""~
water • 25001Imin.

E Inner face of the mold


{E. 200
\._--~~
~-o.-o..
... ~i~
100
...
... ,...... '.,:-
..........................
'
............
I
-=i ==-=. ':.::.~:: :::::. - - _
.

Outer face of the ;,~ict::::.-=.-:.-=i
O~------~-----~----~-------~-----~--~
o 100 200 300 400 500 600

Distance from break ring (m m)

Fig. 8 Temperature distribution in the mold for HCC of 115 nun


square steel billets (from reference [18]).
42

-:c x 106
(\IE
" 10.-----------------------.
8
mold material o - cup
8 i mold length
( mold thickness
mold water
600mm
12mm
2100.GAnin
o

-
\
6
c

-e
Q)

-- 2.5 m/min cOnventlonal slab


4
j
L.
Q) \. (1.4 m/min)
U)
c
e
-
c
Q)
.I::.
2
-----
0 _____

-------::~~~-
'--...::..~-.----
8 =- 0::-
.3
o 0 100 200 300 400 500 600
Distance from break ri ng (mm)

Fig. 9 The local heat transfer rate (i.e., flux) in the mold
(from reference [18]). The curve for the conventional
slab is for vcc.
43

2.2 Thermophysical Properties


2.2.1 Heat Transfer Coefficients
For modelling HCC of ferrous alloys, where the
temperature distribution in the mold assembly itself is
important, it is necessary to accurately estimate heat
tr.ansfer coefficients. Heat transfer coefficients collected
from the literature are summarized in Tables 2 through 5 and
are discussed below.

2.2.1.1 Heat Transfer in the Mold-Metal Gap


Gap heat transfer coefficients (hgOp ) for various
continuous casting scenarios are given in Table 2, where the
results, except those of Adenis et ale [30] and Michalek and
Dantzig [35], are based upon calculations that assume a gap
thickness.
For the casting of a 0.4 m dia. billet of a Mg alloy
(9.5% AI, 0.5% Zn, 0.2% Mn), Adenis et ale [30] give radial

heat fluxes and measured surface temperatures from which the


values of hgOp in Table 2 were calculated herein. These
calculations are described in detail in 5.2.1.1.
The local heat extraction rate through the gap for HCC of
steel shown in Fig. 9 were used to calculate hoop. Because of
the contraction of the solidifying shell away from the mold
face, the heat flux falls drastically from 191 cal cm- 2 s-1 at
the mold inlet to about 82 cal cm- 2 s-1 only 5 cm downstream.
44
Table 2
Heat Transfer Coefficients in the Mold-Metal Gap

Heat Transfer
Coefficients
Application cal cm"2 s"' C·,
0 Remarks
Mg-8% Al (VCC)
Adenis et al. [30] 0.017 to 0.013 Calculated herein
from measured heat
fluxes
steel (VCC)
Michalek and 0.072 to 0.024 Calculated with
Dantzig [35] model that
includes thermal
distortion.
steel (HCC)
Miyashita et al. 0.172 to 0.029 Calculated herein
[18] using Figs. 9 and
10.
steel (HCC)
. Pehlke and Ho [41] 0.072 to 0.047 Assumed to be
uniform and equal
to chill casting
data.
Electroslag Refining
Yu [47] 0.145 to 0.005 Calculated with
simple model.

such heat fluxes are 50 to 300 percent higher than measured


values for slab casting by vee at 1.4 m min"'. It is worth-
while to note that heat fluxes reported by Miyashita et al.
[18] agree reasonably well with those reported by Singh and
Blazek [43] and Blazek [44-46]. These latter workers did not
measure billet temperatures; therefore they could not extend
their analysis to the calculation of hgap.
45

Heat transfer across a gap depends on both conduction and


radiation through the gaseous medium in it. The gap heat
transfer coefficient (also called the effective heat transfer
coefficient by some authors) is

(2.1)

where kgap is the thermal conductivity of the gas in the gap


and 0gap is the thickness of the gap.
Based on previously measured heat fluxes in molds used
for casting steel billets by vee, Brimacombe et ale (32]
calculated the distortion along the hot face of the mold and
found that the mold bulged out near its hottest zone, where
the gap was estimated to be about 0.6 mm. However, they did
not attempt to calculate the gap heat transfer coefficient.
In a more complete analysis, the gap thickness and gap heat
transfer coefficient for vee of steel slabs were obtained by
Michalek and Dantzig [35] via numerical iteration using the
commercial ANSYS finite element package. Beginning with an
assumed gap heat transfer coefficient, the temperature in both
the metal and mold was calculated, and based on the tempera-
ture field the distortions in both casting and mold were
computed. The calculations were repeated until hoap at each
location along the length of the mold converged.
The heat transfer coefficients estimated by Michalek and
Dantzig [35] are shown in Fig. 10. At the meniscus, the melt
46

is in contact with the mold, and the gap thickness increases


to 0.13 rom at 400 rom below the meniscus. The corresponding
values of hoap range from 3.0 x 10-3 W·mm- 2 • °C-' (0.0717 cal cm- 2
s-' °C-') at the meniscus to 1.0 x 10-3 W·mm- 2 .oC-' (0.024 cal cm- 2
s-' °C-') 40 cm below the meniscus. Assuming a reasonable
temperature difference of 1000°C between the shell and the
copper mold, these hoap would result in local heat extraction
rates similar to those reported by Singh and Blazek [43] and
Blazek [44-46]. An interesting result showed by Michalek and
Dantzig was that the radiation heat transfer in the gap was
only about 6% of the total even at casting surface temperature
of 1400°C and mold surface temperature of 300°C. Therefore
considering only conduction heat transfer is an acceptable
approximation for heat transfer in the mold-metal gap.
Yu [47] has calculated from first principles the overall
heat transfer coefficient in the electroslag refining process.
For an ingot surface temperature of 1260°C and a crucible
surface temperature of 95°C, the radiation contribution to the
overall heat transfer through the ingot-crucible gap was only
170 J m2 K-' s-' (0.41 X 10- 3 cal cm- 2 s-' °C-') while the conduc-

tion contribution varied from 6000 J m- 2 K-' s-' (0.143 cal cm- 2
s-' ° C-,) for a gap width of 0.01 rom to 200 J m- 2 • K"' s-' (0.005
cal cm- 2 s-' °C-') when the gap width was 0.2 rnm. These values
also confirm Michalek and Dantzig's [35] numerically iterated
gap heat transfer coefficients.
0.003 0.2
\.
Contact
Region

.
.,...
0.002 E
0° E
C}I (/)
E (/)
CD
E 0.1 12
~ u
~
a.
ro a.
OJ ro
.r:; 0.001 (!)

0.000 I I i I 0.0
o 100 200 300 400
Distance from Meniscus mm
I

Fig. 10 computed gap thickness and heat transfer coefficient for VCC
of 82 mm dia round billet of 0.7% C steel at 21.16 mm s·1 [35].
,c:..
-..I
48
In their work on simulating HCC of steel, Pehlke and Ho
[41] assumed a uniform gap heat transfer coefficient along the
entire length of the mold. For the stationary part of the
withdrawal cycle they used 3000 W·m- 2 • o C- 1 (0.072 cal cm- 2 s-1
OC- 1) and for the forward stroke a value of 2000 W.m- 2 • o c- 1
(0.047 cal cm- 2 s-1 OC- 1) was used. similar range of values for
hgap were used by Croft et al. [39] though the distribution
along the mold or other details were not given. Taylor and
Zalner [26] also estimated the thickness of the solidified
shell next to the mold in the HCC of steel; however, no gap
heat transfer coefficients were given. Neither could the gap
heat transfer coefficient during HCC of steel be estimated
directly from the work of Miyashita et al. [18]. Therefore,
an alternative procedure was followed. The gap heat transfer
coefficient for VCC of steel slabs found by Michalek and
Dantzig [35] (Fig. 10) were scaled up by the ratio between the
heat fluxes of the VCC and HCC processes (Fig. 9), and the
result is shown in Fig. 11.

2.2.1.2 Mold-Water Heat Transfer


Various mold-water heat transfer coefficients from the
literature are shown in Table 3.
Adenis et al. [30] used a constant heat transfer
coefficient of 4 x 103 Btu ft- 2 hr- 1 a F- 1 (0.54 cal cm- 2 s-1 a C- 1)
for their modelling of direct-chill casting of Mg-8% Al
49

Table 3
Heat Transfer Coefficients for Mold-cooling Water

Heat Transfer
Coefficients
Application cal cm- 2 S-1 C- 1 0 Remarks
Mg-S% Al (VCC)
Adenis et ale [30] 0.54 By numerical
iteration to
fit measured
surface
temperatures.
AI-0.6% Mg (VCC)
Weckmann and 0.5 By numerical
Niessen [34] iteration.
steel (VCe)
Dippenaar et ale [32] 0.4-1.03 Calculations
based on heat
transfer
correlations.
Michalek and LOS
Dantzig [35]
Steel (HCe)
Miyashita et ale [IS] 1. 2-0. 55 Calculated in
this work from
empirical data
[lS].

alloys. However, mold cooling may include both forced


convection as well as nucleate boiling with its much higher
heat transfer.
Samarasekera and Brimacombe [7] used heat transfer
correlations for such mixed cases. They found that the heat
transfer coefficient for forced convection is only a weak
function of temperature, because the properties of the fluid
depend on temperature. On the other hand, under conditions of
50

0.2 , . - - - - - - - - - - - - - - - - - - - - - - .

-b
...
~ vee, Michalak & Dantzig [35]
....
0

'(Jl
0.1 • Hee , scalad from Miyashita at. al [18]
~
E
0 .
(ij
0
'"--"'
a.
R!
OJ
.r:::.

0.0 -t--..,....-..,.---r---r---r--,.--,..--r---r----1
o 10 20 30 40 50

Distance from Meniscus (cm)

Fig. 11 Estimated gap heat transfer coefficients for HCC of


steel billets. The value of hoop given by the broken
line on the upper curve is assumed to be constant as
s,;!ggested by hoop in the "contact region" shown in
F~g. 10.
51

nucleate boiling the heat-transfer coefficient increases


substantially with mold-wall temperature. similar calcula-
tions for nucleate boiling as well as the boiling flux were
done by Weckmann and Niessen [34] in heat transfer
calculations on the VCC of aluminum alloys.

By measuring the temperature rise of the mold cooling

water in HCC, Miyashita et ale [18] calculated the average


heat flux as a function of the casting velocity. Using the
measured temperatures in the mold given in Fig. 8, and mold
cooling water temperature profile obtained from the heat flux
data of Fig. 9, an estimation was made of the heat transfer
coefficient for water cooling of the mold (Fig. 12) which
ranged from 0.55 to 1.43 cal cm- 2 s-1 oc- 1 • Note that the heat

transfer coefficient goes through a minimum just upstream of


the inlet. The minimum may be caused by transient heating of
the mold adjacent to the break-ring during the pull phase of

the withdrawal cycle (discussed in §6.4.1 in more detail).

2.2.1.3 Secondary Cooling Heat Transfer

Heat transfer coefficients for spray cooling of the


casting downstream of the mold are presented in Table 4.

Adenis et ale [30] used a constant submold heat transfer


coefficient of 2 x 103 Btu ft2 hr- 1 °F- 1 (0.27 cal cm- 2 s-1 OC- 1)

in their model for casting of magnesium alloys.


160 ,-----------------------------------------------~1.4

140
--0- Mold Outer Face Temperature
1.2 .-
. 0
0

0
C)
120 Mold - Water Heat Transfer Coefficient
.
.-
II)
1.0 ~
!!!
:l §
e 100
8. B
E
~
0.8
~
80 .c

0.6
60

40 -1--......-----,.-----.----.---.---.--...--......-----,.----,.----.---+1 0.4
0 10 20 30 40 50 60
Distance from break-ring. cm

Fig. 12 Calculated local heat transfer coefficient for water-cooling of HCC mold
based on data of Miyashita et ale [18].
UI
N
53
Since both pure Zn and Al-Mg alloy have relatively low
solidification temperatures, it is possible that radiation and
film boiling common for cooling ferrous alloys did not play a
part in DC casting of these alloys reported by Weckmann et

Table 4
Heat Transfer Coefficients for Spray Cooling

Heat Transfer
Coefficients
Application cal cm"2 s"' • C"' Remarks
Mg-8% Al (VCC)
Adenis et al. [30] 0.27 Numerical
iteration to fit
temperatures in
casting.
Zinc
Weckmann et al. [33]
water jet 1.3 Numerical
iteration.
water film 0.86 to 0.29 Same. Higher
value just below
mold.
Al-0.6% Mg
Weckmann and 1.68 to 0.36 Numerical
Niessen [34] iteration.
Steel (VCC)
Miziker [31]
water jet 0.05 to 0.01 Assumed.
Larrecq et ale [49] 0.005 to 0.019 Experimental.
Spray Cooling
of Copper (S800·C)
Bamberger
and Prinz [50] 0.011 to 0.029 Experimental.
Quench Steel
in stirred Water
Lambert and
Economopoulos [51] 0.01 Experimental.
54
Table 4
Heat Transfer Coefficients for Spray cooling (continued)

Heat Transfer
Coefficients
Application cal cm- 2 s-' • C-, Remarks
Steel (HCC)
this work 0.005 From correlation
of Ref. 49 using
secondary cooling
water flow rate
of Ref. 18
0.002 to 0.006 By extrapolating
data of Ref. 50
to 1050·C and
flow rate of
Ref. 18

al. [33,34]. By defining the thermal field during D.C.


casting of zinc billets [33] via thermocouple measurements,
heat transfer coefficients for various cooling configurations
were determined. Directly under impinging water jets below
the mold the value of h was 5.5 W cm- 2 ·C", (1.31 cal cm"2 s"1
DC-') and 3.6 to 1.2 W cm"2 (0.86 to 0.29 cal cm"2" s"' CO,) for
water film falling down the billet surface and cooling it down
from 100 to 30 C in 30 cm.
D

For the direct chill VCC of Al-Mg alloys into 15 cm dia.


billets, Weckmann and Niessen [34] calculated heat transfer
coefficients for spray cooling based on nucleate boiling with
forced convection. They used these calculated heat transfer
coefficients to estimate the temperature field in a billet
55
cast at 1.69 mmos-'. The predicted shape of the liquid sump
agreed very well with the shape determined experimentally, so
this gave credence to the calculated heat transfer
coefficients.
The heat transfer coefficient decreased from a high of
7 x 104 W m- 2 K-' (1.68 cal cm- 2 s-' °e-') at the mold exit to
0 0

1. 5 X 104 Wom2 K-' (0.36 cal cm- 2 s-' °e-') at 5 cm below it.
0

In modelling heat transfer during vee of steel slabs,


Miziker [31] used spray heat transfer coefficients ranging
from 65 to 400 Btu ft- 2 0 F-' hr- 1 (0.01 to 0.05 cal cm- 2 s-' °e- 1 )
with the stronger sprays being used closer to the mold exit.
Miziker's choice of heat transfer coefficients was validated
by the experimental data of Auman et al. [48] for cooling
horizontal surfaces, at temperatures up to 2000°F {1093°e),
with a fan type spray of water.
More recently Larrecq et al. published heat transfer
coefficients based on industrial data for vee of steel billets
and slabs [49]. The surface temperatures of billets and slabs
were measured at the entrance and the exit from the cooling
zones. The surface temperatures thus obtained were used as
boundary conditions in a heat transfer solidification model of
continuous casting, and the heat transfer coefficients were
calculated therefrom. Their results are shown in Fig. 13,
where it can be seen that the heat transfer coefficient is a
strong function of the specific water flow rate.
56

Bamberger and Prinz [50] carried out a controlled set of


experiments on spray cooling of copper billets. The billets
were inductively heated up to 700·C and cooled by water
sprays. Imbedded thermocouples provided temperature versus
time in the billets, and from the cooling curves they were
able to determine heat transfer coefficients as function of
specific spray water flow rate and billet surface temperature.
Their results are shown in Fig. 14; the data are useful for
predicting the heat transfer coefficients in HCC of billets,
in general. Bamberger and Prinz [50] also provided a semi-
empirical equation based on Fig. 14; their equation for the
spray heat transfer coefficient hspray (in W m- 2 K- 1) is:

~spray - 0.69 log [ Vw


0.0006
) [1.4 Vk P C exp rO.32 Os -
l 0b-Oe
Oe) + hY )

(2.2)

The terms in Eq. 2.2 are explained in the List of Symbols. If


surface temperature of the billet exceeds 700°C then hy takes
a value of 750 W m- 2 K- 1• This equation is of course valid
only for water spray flow rates exceeding 0.6 X 10-3 m3 m- 2 s-1.
The experimental results of Larrecq et ale [49], Fig. 13,
and those of Bamberger and Prinz [50], Fig. 14, were used to
estimate appropriate heat transfer coefficients for the HCC of
115 mm square steel billets reported by Miyashita et ale (18].
For the specific water flow rates of only 0.15 x 10- 3 to 0.4 X

10-3 m S-1 (9 to 24 liter m- 2 min-') employed in this example of


57

IJOO

i' 1000

I ... '- +~
+.... l

100
. . ./ ~/
.........+

P?: .., ~+ .........

'0 0
./. .
20 0
// V"""'"

o L7S SoS &U 11

SpecIfic wltlr !low 'It I (10-' m'/m'/sl

,. SQUARE BlUETS '20 • 120 (mm'J


2. ROUND BII.LETS 0 = '20 mm

Fig. 13 Heat transfer coefficients in the secondary cooling


region [49]: measurements carried out on a billet
caster.
58

10Sr-----r-----~----~----------

t-"
z
w
u
u::
u.
w
o
u
a:::
lJJ
u.
Vl
Z
<!
a:::
t-

~
lJJ
:c

Fig. 14 Heat transfer coefficient as function of billet


surface temperature during spray cooling at various
spray intensities [50].
59
HCC, from Fig. 13 a hspray value of 200 to 220 W m- 2 K-' (about
0.005 cal cm- 2 s-' °C-') was determined. For the same flow rates
Fig. 13 indicates somewhat lower values of hspray for round
billets, namely 100 to 130 W m- 2 K-' (about 0.003 cal cm- 2 s-'
It is noteworthy that by extrapolating the data of
Bamberger and Prinz (50], Fig. 14, to the specific water flow
rate of 9 to 24 liter m- 2 min-' used in HCC of steel (lS] and
assuming a billet surface temperature of 1050 ° C, hspray of
190 W m- 2 K-' (0.0045 cal cm- 2 s-' °C-') was obtained, indicating
close agreement with the values obtained from results of
Larrecq et ale (49].

2.2.1.4 Air Cooling


Immediately downstream of the mold and beyond the
secondary water cooling section (s), the billet cools by
radiation and natural convection. Although the heat transfer
phenomena for this case are simpler than those discussed above
and more amenable to estimating, it is still preferable to
have experimental data. Lambert and Economopoulos (51]
experimentally determined the heat transfer coefficients for
cooling of steel sheets in still air for temperatures as high
as SOO°C. The heat transfer coefficient is 0.004 cal cm- 2 s-'
°C-, at SOOGC, comparable to the value by spray cooling at that
temperature for HCC (Table 4). Availability of such empirical
data removes the need for using non-linear boundary conditions
60
due to Stefan-Boltzman equation in subsequent mathematical
modelling.

2.2.1.5 Heat Transfer in the Melt-Pool


The thermal conductivity of the melt used in various
simulation of continuous casting processes is given in
Table 5. Heat transfer in the melt pool in a typical con-
tinuous casting process may be enhanced by natural or forced
convection. The additional heat transfer due to this convec-
tion have been taken into account in several ways by various
workers.
Table 5
Thermal Conductivity in the Melt Pool

Thermal Conduc-
tivitv
Application cal cm- 1 s-1 c- 1
0 Remarks
Mg-8~ Al (VCC)
Adenis et ale [30] 0.1034 Convection ef-
fects neglected.
Al-0.6% Mg (VCC)
Weckmann and - Same as pure Al
Niessen [34] at the melting
point.
steel (VCC)
Miziker [31] 0.62 7 times that of
stagnant liquid.
Michalek and 0.1155 to 0.175 3.5 times
Dantzig [35] higher.
Dippenaar et ale [32] - Convection
neglected.
61

In his model for the vee of 6 in. thick steel slabs,


Miziker used a thermal conductivity of the melt pool that was
seven times higher than that for stagnant liquid steel.
Miziker obtained this value of thermal conductivity by
matching model calculations to experimental solidification
rates in static castings. For their simulation of vee of
steel, Michalek and Dantzig [35] chose to raise the thermal
conductivity of the melt by an arbitrary factor. The liquid
thermal conductivity was linearly increased from its nominal
value at the liquidus temperature to a value 3.5 times that
value at 100 e above the liquidus.
0
Dippenaar et al. [32] on
the other hand did not take into account any effect of
convection on enhancing the apparent conduction in the melt
pool for modelling vee of 14.4 cm sq. steel billets.
Thomas [52] has reported a more rigorous treatment for
vee, in which the recirculatory flow in the melt pool is
solved along with the thermal field. Therefore unlike the
above examples no arbitrary assumptions had to be made
regarding the effect of convection in the melt pool. However,
no ready estimate for the effective thermal conductivity in
the melt pool could be made from the results of this work.
None of the thermal models for Hee [38-41] published to date
has mentioned the effect of convection in the melt pool
explicitly.
62

The average velocities of flow in the melt pool for HCC


can be expected to be lower than those for VCC, because,
unlike the latter, there is no central core of forced convec-
tion. Therefore, it is probably justified if no adjustments
are made to account for natural convection in the melt pool
for HCC. The departure from a simple plug-flow assumption as
far as heat conduction in the melt pool is concerned will
perhaps be noticeable only in the neighborhood of the break-
ring during periods of acceleration. Convection will of
course affect the complete melt pool in case of HCC if
electromagnetic stirring is used.

c.3 Mathematical Modelling of continuous casting


Well documented published work on mathematical modelling
of continuous casting appears to be confined to steady-state
models and not to transient models necessary for HCC. The
steady-state results were useful to check the steady-state
behavior of the transient model developed herein.

2.3.1 vertical Direct Chill Casting of Mg Alloys


Adenis et ale [30] calculated steady state temperature
profiles for VCC of Mg-8% Al and Mg-5.5% Zn alloys cast into
round billets (0.4 m dia.). The calculated temperatures
agreed quite well (Fig. 15) with temperatures measured by
them.
63

MOULD END

=.:=.......,.:=--"___ CENTRE
----t--- --xl
i ------.J~<~ LIQUIDUS
_.- - _.- -'~--"""" - -j-'-'-'-'-' ":,'-'-'-'-
I "

1000
i '\ ,
\
\
\
\
\

... 900 \ ,,
~
""
S 800 - - -'-'-'-'- -'-'- _.- -'.,
, SOLIDUS

w I
i ~
g
I- 700
.,
II
i ,l
600 II
ji
soo
i
i
PREDICTED
OBSERVED
~oo

Fig. 15 Comparison between calculated (---) and experimental


(-) temperature profiles for VCC of AZ80A alloy;
billet dia. 16 in.; mold length 9.5 in.; casting
speed 2.1 in min- 1 (from [30]).
64
Using the Kirchoff Transformation, they converted the
temperature in the conduction equation to a new variable in
order to remove the non-linearity of the temperature depen-
dence of thermal conductivi ty. The transformed partial
differential equation was discretized by using the finite
difference method (FDM) with a simple mesh refinement. The
FDM equations were solved by the Liebmann iteration technique.
Unknown parameters related to the use of a mold lubricant were
determined by forcing the solved temperature to agree with
measurements, and then the same parameters were used for all
the cases. Heat transfer coefficients in the gap were
calculated by a model that took into account conduction and
radiation through the lubricant film and the air gap. Using
the heat transfer coefficients thus calculated they were able
to simulate the measured temperatures quite closely. However,
the values of their gap heat transfer coefficients in the
lower part of the mold appear somewhat low, probably because
they assumed that the mold temperature was uniform. More
importantly, the assumed mold temperature (370°C) was much
higher than one would expect for casting an alloy with melt
temperature of only 674°C because measured mold temperatures
in casting steel with melt temperature above 1500°C are only
in the range of 50 to 400°C (Fig. 8).
65

2.3.2 vertical Direct Chill Casting of Al Alloys


Weckmann et al. [33] developed an iterative steady state
model for vertical direct-chill (DC) casting using the finite
element method (FEM) with an adaptive mesh for improving
accuracy especially at the liquid-solid interface. They
tested their model against data obtained from laboratory-scale
DC casting of zinc alloys. They used an iterative scheme such
that the converged temperatures had to match measured
temperatures along the billet surface below the mold.
In a subsequent paper, Weckmann and Niessen [34] used the
model for calculating the temperature field in DC casting of
Al-O.6% Mg billets. Unlike the previous case involving the
zinc alloys, billet surface temperatures were not available.
Temperatures were measured at several radial locations, and
the liquid-solid interface was marked chemically with a tracer
added to the melt. Temperature dependent thermophysical
properties of pure aluminum were used' for the simulation.
Because surface temperatures were not available, the cal-
culated surface temperatures in the mold were ad}usted until
a good fit was obtained between the calculated solid-liquid
interface profile and the measured profile. Then, agreement
with the measured temperature field was found, as shown in
Fig. 16.
66

... .., ....


N
...
C>
.,; ...
II> co
on ,.: .,;

THERMOCOUPLES
...
...on

:g LIQUID
iii

E N
H-
'0
'"
...

:z
--
DlD
I-
- 2

NUCLEATE
BOILING
CJ')
D
COOLING
a..

N
...
~

FORCED
CONVECTION
COOLING

...d
N

---INTERFACE BY TRACER
A INTERFACE BY THERMOCOUPLES
--CALCULATED INTERFACE' 923 K)

Fig. 16 comparison between calculated isotherms and measured


solid-liquid interface profile for a 0.1524 m
diameter A6063 aluminum ingot cast at 3.81 X 10.3
m·s· 1 by VCC (from [34J).
67

2.3.3 vee of steel


In both of the above works, the authors had to measure
temperatures in the actual casting in order to numerically
determine the boundary conditions from their respective
models. Bamberger and Prinz [36] took the alternative
approach of using in their steady state model heat transfer
coefficients for secondary cooling sprays by first determining
the coefficients [50] with prototype spray cooling experi-
ments. They used these coefficients to specify boundary
conditions for vee of stainless steel slabs. It is not clear
what water temperature was used to complete the description of
the boundary condition. The steady state POE was solved for
temperature at surface as well as centerline by solving
explicit FO equations by successive over-relaxation. The
model was successfully applied to the vee of Al-2% Mg and the
measured sump profile was matched by calculations. Moreover,
starting with known boundary conditions, Weckmann's tempera-
ture measurements during casting A6063 billets [34] were
reproduced. If the experimental and modelling work by
Bamberger et al is consistent, as it indeed appears from their
papers [36,50], then this latter method of applying known
boundary conditions, computationally more economical for
unsteady state modelling than the iterative method of Adenis
et ale [30] or Weckmann et ale [33,34], appears to be more
suitable for Hee simUlation.
68

2.3.4 HCC of Steel


In their modelling of the HCC process called HORICAST
[18], Croft et al. [39] neglected axial conduction and the
influence of convection in the melt pool. The temperature
fields in the axi-symmetric billets were calculated with an
explicit finite difference method. The authors calculated the
thickness of the mold-metal gap, and the gap heat transfer
coefficient, but very few details were given. Gap heat
transfer coefficients of 4000 to 6000 W m- 2 K- 1 (0.096 to 0.144
cal m- 2 s-1 C- 1) were calculated. No other details of heat
0

transfer coefficients or boundary conditions are available


from this paper. The results included comparisons between
calculated and measured shell-growth, and calculated and
measured mold temperatures. However, the value of this work
is compromised by the lack of detail and the neglecting of
axial heat transfer, at least during the initial shell
formation in HCC.
A similar lack of detail affects the paper describing an
unsteady state model for HCC reported by Holleis et al. [40].
Their model accounted for cyclic withdrawal and reproduced the
pattern of witness marks of ferrous alloys produced by HCC.
An implicit finite difference method with a non-uniform
rectangular mesh was used to solve the thermal field and
solidification thickness near the triple point. They used the
model to study the formation of the gap between the stationary
69

part and the moving part of the shell, i.e. the formation of
a secondary witness mark (Fig. 6). Fluid flow was implicitly
taken into account by the use of an enhanced thermal conduc-
tivity for the liquid. The boundary conditions were derived
from measurements of heat flux in the mold, as well as data on
rapid solidification. Calculated isotherms, at an inter-
mediate time (0.45 s) in a 0.6 s cycle, showed solidification
contours similar to those depicted in Fig. 6. In solving for
the isotherms related to formation of the witness mark, it is
not clear what boundary conditions for the bulk liquid were
used or if the larger scale heat transfer in the whole billet
was solved.
Pehlke and Ho [41] predicted the shell profile in HCC
under several operating conditions by using numerical method-
ology identical to those of Holleis et al. [40]. However,
they claim to have taken into account the effect of fluid flow
in the region upstream of the secondary witness mark. To
model the formation of the witness marks, the shell in this
part of the casting was divided into 3 to 5 segments (each
resulting from a complete cycle) with at least 9 nodes per
segment in the withdrawal direction. At the end of the
solution field (3 to 5 segments downstream of the break-ring)
the longitudinal flux was considered to be zero, and the
temperature of the mold was assumed to be a constant 350°C.
As the overall temperature field was not solved, boundary
70

conditions in other regions were not specified. Nevertheless,


they examined the effects of stroke length and frequency,
break-ring height, and superheat on the solidification of the
newly formed shell. However, no comparison with experimental
shell profiles or other empirical data were given. By not
solving the overall thermal field, the utility and probably
the accuracy of their work were compromised.

2.4 Numerical Methods


2.4.1 Mesh Refinement
For both rectangular and irregular meshes accuracy can be
improved by reducing the mesh spacing in regions of high
gradient in the field variable. Sundquist and Veronis [53]
evaluated the effect of varying the mesh spacing by solving
for the stream function in a boundary layer, for which the
analytical solution was available. With an uniform coarse
mesh, numerical oscillations in the solution were observed.
Smooth and accurate solutions were obtained by reducing the
mesh spacing in an uniform mesh. Mesh refinement near and
within the boundary layer using smoothly varying spacings also
gave satisfactory solutions.
Thacker [54] reviewed various means of refining meshes,
among which is the technique for generating two dimensional
meshes with smoothly varying spacings. A coordinate trans-
formation of an uniform mesh is done, and the technique
71
appears to be both easy and amenable to analysis for stabil-
ity. A coordinate transformation technique applied to
boundary layer problems also has been discussed by sills [55].
Gal-Chen and Somerville [56] used the technique for transform-
ing irregular boundaries into rectangular form while solving
the Navier-Stokes equation.

2.4.2 composites
Layered composites are characterized by internal inter-
faces with different properties on each side. If temperature
is continuous, even though the thermal gradient is not, then
the interface is referred to as a Type I interface (Fig. 17a).
When the temperature is not continuous, as is frequently the
case for the mold-metal gap in castings, there are two
distinct temperatures, one at each side of the interface
(Fig. 17b) , and the heat transfer across the interface can be
characterized by a finite gap heat transfer coefficient. This
latter type of interface is referred to as a Type II
interface.
Special procedures become necessary to obtain the finite
difference equations at interfaces. Carnahan et ale [57]
derived a finite difference expression for one-dimensional
heat transfer across such an interface by combining the Taylor
series expansion of the derivatives in the Partial Differen-
tial Equation (POE) on each side of the interface with the
72

o(O,t) = 01 ' - - -

Type I Interface

k J aD
o(L,t) =0 2
D
a~

L ------II.....~
(a)

Type II interface

o(L,t) = O2
a ~

L -------113_

(b)

Fig. 17 Schematics for one-dimensional heat transfer in two


layer composites with (a) Type I internal interface
having infinite heat transfer coefficient,
(b) Type II internal interface with finite heat
transfer coefficient.
73

continuity of flux on both sides. They did not extend this


analysis to two dimensions or more complicated geometries e.g.
reentrant corners.

2.4.3 Reentrant Corners


If the direction of the boundary of a re~:don changes
suddenly by an angle exceeding ", then the resulting corner
defined by the angle is called a reentrant corner. Singulari-
ties occur at the reentrant corners because some of the
partial derivatives become unbounded. Finite difference
methods, therefore, have questionable validity in the neigh-
borhood of reentrant corners as, even for otherwise stable
algorithms, the numerical errors at such nodes propagate into
the adjacent "regions of infection" [58].
The singularity at a reentrant corner with angle a > ",
may be demonstrated simply by considering Laplace's equation
for the scalar variable, u; it is

(2.3)

in circular cylindrical (r,v) coordinates [59] with the origin


at the corner. If the boundary conditions require u to vanish
at the linear boundaries that meet at the reentrant corner,
then the solution [58] is

iii mr
u(r,v) - L anr a sin (n~v) (2.4)
n-'
74
where the an s are constants. If on the boundaries the normal
derivatives are zero, the following solution results

(2.5)

where the b n s are constants.


For both cases however, when n = 1, the exponent of r in
the above equations is less than 1 if a >~. Therefore, the
derivatives au/ar are unbounded at the origin (r = 0).
There are at least three ways of dealing with
"-
singularities at reentrant corners. These are discussed in
the following sections.

2.4.3.1 series Solution


For the solution of Laplace's equation in a field that
contains a reentrant corner wi th ~ S a S 2~, Motz [59]
proposed a simple method based on the above series solutions
for u. At most nodes in the field, a formulation obtained
from a standard five point finite difference formulation is
retained. But for the nodes immediately surrounding the
singularity, the finite difference approximations are
replaced, depending on the boundary conditions, by one or the
other of the above series solutions (Eq. (2.4) or Eq. (2.5».
The unknown coefficients an (or b n, as the case may be)
are evaluated as follows. When boundary conditions of the
75

derivative type are involved, Eq. (2.5) is appropriate. If


only the first four terms of that series are used, then

u - b o + b, r 1/2 cos ~ + b2 r cos II + ~ r3/ 2 cos 311 (2.6)


2 2

At the arbitrary kth iteration step during the finite


difference solution of the elliptic Laplace's equation, the
coefficients b o b3 are obtained by using the values
(obtained by finite difference) at four remote nodes
(j = 1 ••• 4). Thus four linear equations of the form

(2.7)

result. These are solved for bl , thus obtaining the


coefficients necessary to compute the u values at the nodes
immediately surrounding the singularity. The coefficient b o
provides an acceptable approximation for the value of u at the
singularity.
The accuracy of the series solution may be checked by
applying it to other remote nodes (which were not used for
determining the coefficients) and then comparing the results
to the values calculated for those points by finite
difference. This procedure becomes mandatory when it becomes
desirable to minimize the number of remote nodes used to
determine the coefficients a i or b i • For reentrant corners
with Q = 3~/2, the algorithm may be started with only three
terms in the appropriate series, but more terms and remote
76

nodes are needed if errors become unacceptable. This simple


procedure is attractive when higher precision or wider mesh
spacings are required, because improvement' of the Motz
algorithm does not require more costly iterations over the
whole field, but just an increase in the number of terms in
the series and a recalculation for the nodes surrounding the
singularity.
For unsteady state diffusion in a two dimensional field
containing a reentrant corner, singularities similar to those
in the steady state arise. Bell and Crank [60] extended
Motz's method by first rewriting the POE in polar coordinates
at the singularity and then derived series solutions consist-
ing of Bessel functions. For a reentrant corner with ex = 31f/2

and linear boundaries on which the normal derivatives are


zero, they obtained

CD CD

u(r,v,t) - EE Akj exp (-A~t)J2k (Aj r) cos (2~V)


j-() k-o 3

(2.8)

where AJ are roots of equations containing Bessel functions,


and the AkjS and Cks are constants.
77
If only the first four terms of the series are taken into
account the following relation results

u (r, /I , t) - ao (t) + a, (t) cos ( 2;) r;2!3 + az (t) cos ( ~) r 4!3

+ r2[a3(t) cos 2/1 - bo(t)] + o{rB!3) (2.9)

which is equivalent to Eq. (2.6) obtained by Motz for elliptic


problems. Therefore, the method of Motz can be applied to
reentrant corners in problems of two dimensional transient
diffusion fields.
Bell and Crank [60] found that it was better to have a
large series approximation at the cost of a large number of
treated nodes. In addition, they showed that better accuracy
at the nodes surrounding the reentrant corner is obtained not
only by increasing the number of terms in the series solution,
but by also ensuring that the remote nodes (which are used to
evaluate the series coefficients) are in the proximity of the
reentrant corner, itself.

2.4.3.2 Modification of the Finite Difference Equations


Approximations based on a Taylor series expansion in the
vicinity of the singularity are not accurate, especially when
there are steep gradients of the field variable. If, however,
the derivatives in the PDE can be replaced by a relation
involving values at neighboring nodes without recourse to the
Taylor series, then such inaccuracies are avoided.
78

Instead of the Taylor series, Bell and Crank [61]


differentiated Eq. (2.9) to obtain the derivatives needed to
explicitly solve the two dimensional unsteady state conduction
equation at the nodes adj acent to a reentrant corner wi th
a = 3~/2. with this approach, they derived time independent
equations for the unknown coefficients of the series in terms
of values at four neighboring nodes where the conventional
finite difference formulation was used.
First conventional finite difference equations were used
to calculate temperature at all points in the field. Then the
special relations for the corner node and its four nearest
neighbor nodes in the mesh were used to recalculate values at
only those five nodes. This technique eliminated the computa-
tional work needed to solve for the unknown coefficients of
the series at each time step. ThUS, it is an improvement upon
the Motz series method. Although the authors stated that the
extension of this me.thod to an implicit scheme is straight-
forward, the effect of using the special relations at and near
the corner node on the otherwise unconditional stability of
Crank-Nicolson type implicit or ADI methods [57.] was not
explored.

2.4.3.3 Mesh Refinement near Discontinuities


Ames [58] stated that a common method of effectively
dealing with discontinuities is to diminish their effect by
79

using mesh refinement in the region near the singularity.


Trutt et ale [62] attempted the use of a very fine rectangular
mesh in the vicinity of a reentrant corner and iteratively
solved the nonlinear elliptic equation for a scalar potential.
However, their success in limiting the "region of infection"
could have been due to averaging the magnetic permeabilities
across the interfaces.

2.4.4 Solidification
Numerical algorithms developed for solving heat transfer
in solidification involving the evolution of latent heat may
be classified into two categories:
(a) Those that explicitly describe latent heat evolution
at the phase-change boundary and its movement using
the Stefan condition;
(b) Those that lump the latent heat evolution with the
general heat transfer using integral expressions for
enthalpy and then determine the location of the
moving boundary from the enthalpy field thus
obtained.
For modelling the unsteady state thermal field in HCC and the
solution of that model by an implicit FDM, it was necessary to
adopt the second alternative. The following discussion
supports that choice.
80

2.4.4.1 Explicit stefan Condition


Consider one dimensional solidification of a pure
element. The moving interface between the two phases is
described by the Stefan condition:

pt.H( ds _ [-k ~] - [-k ~] (2.10)


dt ax solid ax liquid
where p is the mass density of the solid, t.H( is the heat of
solidification, ds/dt is the speed of the interface, k is
thermal conductivity, and as/ax is the temperature gradient.
When solving for the temperature field, the solution method
must include determining the location of the interface and
selecting the phase-dependent properties accordingly.
Discretization of the governing equations by fini te
difference can be carried out by one of three techniques: the
fixed mesh, variable time step, and variable mesh spacing
[63].

with a fixed mesh, the interface most likely lies between


two nodes, and the temperature derivatives at these two
adjacent nodes are described by an interpolation. An explicit
solution is straightforward, but an implicit solution requires
iteration because the boundary position at the new time-step
is unknown. Moreover, singularities may arise when the moving
interface is too near a node. Hence, there have been attempts
to modify the mesh so that the moving interface always remains
on a node.
81

Rather than using a fixed time step and searching for the
interface, a variable time step is defined such that the
moving interface coincides with the mesh line at each time.
starting with an assumed value of the timestep, the
discretized form of the temperature field is solved with the
invariant melting point as a boundary condition. An iteration
for the unknown timestep is performed, and then by using that
timestep the new temperature distribution is determined.
Unfortunately, there is no obvious way of applying this
technique to two or three dimensional heat conduction.
The mesh spacing can be adjusted from one timestep to the
next such that the moving interface always stays on a set of
,given nodes. This requires a coordinate transformation to
account for the variable mesh spacing. This method has been
extended to cylindrical geometry, but using it in conjunction
with a mesh refinement is computationally laborious.

2.4.4.2 Lumped Latent Heat


During solidification, the solid liquid interface is a
moving boundary that may have sharp peaks, or it may double
back or even disappear (single phase). The reformulation of
the solidification problem in such a way that the Stefan
condition is incorporated within a new form of the energy
equation, which applies everywhere in a fixed domain, appears
attractive. In the enthalpy method the posi tion of the
82

interface is determined a posteriori from the numerical


solution.
In order to get around the discontinuity of the enthalpy
at the interface for pure metals, Shamsundar and Sparrow [64]
originated the following integro-differential equation.

d~ Jv HdV JHv • d A - Jk
+
A A
V 0• dA (2.11)

where H is the enthalpy, v is velocity, A is surface area, and


V is volume. Solving the above enthalpy balance equation with
suitably modified boundary conditions allows the calculation
of temperature from enthalpy.
Furzeland [65] compared the results of solving two-phase
moving boundary freezing problems by two of the discretization
methods discussed in § 2.4.4. 1 as well as by the enthalpy
method and concluded that the enthalpy method was the most
suitable for "mushy phase change problems" or for multi-
dimensional problems with complicated shapes.
For alloys both Hand 0 are continuous and differentiable
so it is not necessary to use Eq. (2.11); rather the usual PDE
form of the energy equation can be used.
83

CHAPTER 3

DEVELOPMENT OF NUMERICAL TECHNIQUES

The temperature field in horizontal continuous casting is


characterized by regions of high temperature gradients,
localized evolution of heat of freezing, imbedded interfaces,
reentrant corners and complex boundary conditions. Moreover,
the need to simulate the effects of withdrawal cycles requires
an accurate solution of the transient field. In order to
solve this problem wi thin the restrictions of a personal
computer, it was necessary to adapt and develop efficient
numerical techniques and strategies. The alternating direc-
tion implicit (ADI) method [66] was chosen for solving the
transient problem because of its inherent stability and speed
and memory efficiency, despite its limitations regarding
irregular geometries, adaptive meshing, etc. when compared to
methods based on triangular meshes [67,68].
In order to achieve accurate results near imbedded
interfaces (Fig. 17) with temperature discontinuities, various
mesh refinement schemes were developed. This was done, rather
than simply increasing the number of nodes, because of
computer hardware limitations. Mesh refinement, if applied
inappropriately, may introduce errors so it was necessary to
investigate various techniques. The method of refinement by
84

coordinate transformation with backward differencing was


selected. Instead of solving iteratively the temperature
fields separated by interfaces [35], the finite difference
discretization technique was modified so that the entire field
could be solved in just one pass. The non-iterative approach
is critical to computational efficiency. Finite difference
procedures for handling reentrant corners were developed as
well.

3.1 Mesh Refinement


The success of the numerical techniques developed for
imbedded interfaces depend on optimum mesh refinement. In
order to study the effect of mesh refinement, the following
test problem for unidirectional heat conduction in a single
piece of material was solved analytically and by two
techniques of mesh refinement.

3.1.1 Test Problem


The POE for the test problem is:

(3.1)

with
8(x,0) = 1.0 for °S x S L (3.2a)

and
8(0,t) = 80 = 1.0, O(L,t) = 8, = ° for t ~ ° (3. 2b, c)
85

where 0t is ao/at and Oxx is a2 o/ax2 • Equations (3.1) and


(3.2a,b,c) were solved by analytical and implicit finite
difference methods using various mesh refinements. Finite
difference methods were evaluated by comparing those results
with the analytical solution.
The analytical solution of this problem is readily
available [69]. It is

to

o(x, t) - 1 - x + ~ L (_~)n exp [-a( n;r t ] sin (n~x) (3.3)


I. 7T n-1

The test problem was solved numerically for L = 1 and


three values of a, namely 1.0, 0.1 and 0.01. In the case of
mesh refinement, the mesh was made coarse near x = 0 and
progressively finer as the temperature gradient becomes
steeper near x = L = 1.0.

3.1.2 Methods of Mesh Refinement


Two methods of mesh refinement were studied:
(i) That of Sundquist and Veronis [53] in which the
finite difference equations are written for a
variable mesh spacing, and
86

(ii) Coordinate transformation of the x coordinate in a


nonuniform refined mesh to the X coordinate in a
mesh with uniform spacings. The procedure for
making this coordinate transformation is given in
the following section (§3.1.3) and Appendix A.

3.1.3 PDE and FDE after Coordinate Transformation


Relationships between the gradients and time derivatives
in the x domain and the corresponding derivatives in the X
domain are required. The thermal gradient is transformed as
follows:

2!
8x
_(2!) (dX)
8X dx
(3.5)

2
8 0 8 (8 0 ) 8 (8 0 / dX) / dx
oxx - 8x2 - 8x 8x - 8 X 8X dX dX

8 20dx 80 d 2x
2
8X dX 8X dX2
(1/~~)
(~~t
2 2
8 0 / (dXr - 80 d x / (dXY (3.6)
8X2 dX 8X dX2 dX
87

As can be seen in Appendix A, values of dx/dX and d 2x/dX2


depend on position in the mesh. In the transformed (X,t)
coordinate, Eq. (3.1) then becomes:

(3.7)

where

80
Ox - ax'

a, - -a d
dX2
2
x I( dX
dX)3

and

The finite difference form of Eq. (3.7) may be written as

n+1 n+1 n+1 n


+ b3 ° i+1 = ° i (3.8)

where
88
and

a2~t
(~X)2

Equation (3.8) can be solved by the familiar Gaussian elimi-


nation of the resulting tridiagonal matrix [57].

3.1.4 Results and Discussion


The test case described in §3.1.1 was solved numerically
using the mesh refinement and coordinate transformation
described in §3.1.3.
starting with the base case of an uniform mesh, several
meshes with increasing refinement near x = 1. 0 (where the
temperature gradient becomes steep) were used. Both 0 and Ox
were calculated by the analytical and the finite difference
techniques. Typical computed temperature profiles for
a = 0.01 using both uniform and nonuniform refined meshes are
shown in Fig. 18. with the uniform mesh at x = 0.9 and
t = 0.4, the temperature computed by the implicit finite
difference method was 0.7202, and the temperature calculated
with the analytical result, Eg. (3.3), was 0.7364. Mesh
refinement by coordinate transformation so that at x = 0.896,
l\x was 0.0505 gave a temperature of 0.7538 compared to
analytical result of 0.7551.
89

(a)

DISIAItCE

(b)

'.7
III
It
:::I '.e
to
C '.1
It
III
Ii. •.•
z:
1/1 •• ,
to

...~--~--~--~--~--~--~--~--~--~~
t.M .... I." •. It .... '.11 '.U •. ,. .... ..,.
DISTAItCE

Fig. 18 Computed temperature profiles for the prototype


problem for mesh refinement. Profiles plotted at
time intervals of 0.2 starting at t = o.
(a) Uniform mesh: at x = 0.9, 6X = 0.1;
(b) Nonuniform mesh with coordinate transformation:
at x = 0.896, 6X = 0.0505.
90
The error in the numerical results with respect to
the analytical results were calculated by the following
equation:

% Error _ Numerical - Analytical x 100 (3.9)


Analytical

In Fig. 19a the effect of Ox at x = 1 on the magnitude of the


error in ox' computed by the two finite difference techniques
(Sundquist and Veronis and coordinate transformation) using
uniform as well as variously refined meshes (described in
Table 4), is shown. The exact value of Ox was calculated from
Eg. (3.4). For the uniform mesh (Mesh 1) the error in Ox is
very large when the gradient exceeds 3.0. Only a moderate
mesh refinement (Mesh 5) substantially decreases the errors in
the gradients calculated by both the Sundquist and veronis
(S&V) and coordinate transformation (C.T.) techniques.
Clearly mesh refinement in the region of steep gradients
improves the accuracy of the numerically determined results.
However, notice that there is an indication that, when using
a coordinate transformation, errors increase when the mesh is
further refined from Mesh 6 to Mesh 7. This unexpec::ted result
indicates the danger of using too much mesh refinement.
Figure 19b shows the effect of an arbitrarily defined
Mesh Refinement Factor (MRF), on the errors in Ox at x = 1,
where

MRF _ 6XLD'lfform
(3.10)
6X re fined
Fig. 19 Effect of mesh refinement on percent error in
temperature gradient at x = 1 computed by implicit
finite difference method: (a) as a function of the
exact temperature gradient: (b) as a function of the
Mesh Refinement Factor. C.T. refers to coordinate
transformation and S&V refers to the method of
sundquist and Veronis [53].
Table 6
Characteristics of Meshes Used to Evaluate Mesh Refinement Techniques

at X = 0.9
Mesh
No. Transformation Equation x Ax MRF*

1 x = X 0.9 0.1 1.0


5 x = 2.0X - 1.5X2 + 0.5X3 0.9495 0.0505 1.98
6 x = 2. 5X - 2. 25X2 + O. 75X3 0.9743 0.0257 3.89
7 x = 2.75X - 2.625X2 + 0.875X3 0.9866 0.0134 7.46
--~ ---

*Mesh Refinement Factor, MRF, defined by Eq. (3.10).

\0
tI)
93

The numerator in Eg. (3.10) is the uniform mesh spacing, and


the denominator is the mesh spacing nearest x = 1 for the
nonuniform spacing. When the errors are shown in this manner,
the increase in the error on going from Mesh 6 to Mesh 7 for
the method involving a coordinate transform is clearly
evident. For a steep gradient, mesh refinement is indispensa-
ble. Notice that with the proper choice of mesh refinement
(Mesh 6), the errors resulting from the coordinate transforma-
tion technique are comparable to or less than the errors that
result from the method of Sundquist and Veronis.
The effect of mesh refinement on the er~or in computed
temperatures is plotted in Fig. 20. Again, when the mesh is
ever refined (Mesh 7), the errors resulting from the use of
the coordinate transformation (Le., C.T.) are larger than the
errors from the method of sundquist and veronis. with all
other meshes the C.T. errors are smaller than the errors from
the S&V solution.

3.1.5 Conclusion
From the above it is seen that mesh refinement is
beneficial for reducing errors in regions of high gradients,
but it needs to be used carefully because the errors can
increase if it is improperly used. For optimally refined
meshes, however, the method of coordinate transformation
94

6
METHOD 9y.

5 • C.T. 1.1

tr;
• C.T. 2.8
g II C.T. 8.9
a:
w 4
::l! --0'- S&V 1.1
0
u.
a
w
=> 3
--0- s&v 2.8
..J --D- S&V 8.9
:;
~
=>
...J ~~- ... --.::t'------- ---~----
a 2
CJ)
CD
«
1

o~----~----~----~~--~~----~----~----~----~
o 2 4 6 8
MESH REANEMENT FAClOR

Fig. 20 Effect of mesh refinement on percent error in


temperature computed by implicit finite difference
as a function of the Mesh Refinement Factor. C.T.
and S&V are defined under Fig. 19.
95

yields lower errors than by the S&V technique. For this


reason, the former method was adopted for subsequent work.

3.2 Heat Transfer across a Type II Interface


In the case of a Type II interface perpendicular to the
x axis (Fig. 17), there are two unknown temperatures 0fC and
0fO. At the interface the following conditions, based on the
original (x,t) coordinates hold:

-kc Ox Ic - -ko I
Ox D' (3.11)

and

(3.12)

where kc and ko are the thermal conductivi ties in regions C and


D, respectively, and llco is the gap heat transfer coefficient.
After transformation to the (X,t) coordinates, the
interface conditions are

(3.13)

and

(3.14)
96

where x' = dx/dX. Equations (3.13) and (3.14) are used later
to obtain the FOEs for the interface problem and to eliminate
the unknown interface temperature, 8 10 , with a backward
difference at the interface, respectively. Likewise on the
O-side

(3.15)

Eq. (3.15) is used to eliminate 8 10 when forward difference is


applied at the interface.
At steady state the temperature profiles on both sides of
the interface are linear, and the interface temperatures can
be easily derived from the interface conditions. They are

de 8, + do 82
1 + do
------::--
1 + d _ 1
, (3.16)
e 1 + do

and

(3.17)

where 8, and 82 are shown in Fig. 17b, and we define interface


coefficients

(3.18)
97

and

(3.19)

3.2.1 Finite Oifference Formulations at a Type II Interface


The FOE for the interface point may be obtained by at
least two procedures:
(1) Consider the FOE for the point ic in Fig. 21 with
accuracy of O[~t,(~X)2]. The FOE involves the
unknown temperatures 0 iC-1' 0 iC' 0 fD+1 while eliminating
the unknown 0 fD by the above mentioned interface
condi tions with forward or backward differencing
involving accuracy of O[~X]. For each side of the
interface 0xx is eliminated from the corresponding
POE by using Taylor's series and obtaining an
expression for Ox' which is then substituted into
the first interface condition to obtain the FOE at
the interface. This equation contains the unknown
0iD' which is eliminated by the first order (a) back-
ward or (b) forward approximation.
(2) The second method involves two FOEs, one each for °iC
and °iDe These equations contain terms for the
temperatures at iC-1, ic, iC+1 (imaginary) and iO-1
(imaginary), iO, iO+1, respectively. Both of the
98

,
\
\ 81C + 1
b
1\
@

k (Xc
c'

x=o a L

Fig. 21 Schematic for central difference scheme with


imaginary points iC+1 and iD-l used to solve
unsteady state heat transfer in a two-layer
composite with an internal Type II interface.
99

FOEs have an accuracy of O[At,(AX)2], and the


temperatures at the imaginary points are eliminated
by relations with accuracy of O[(AX)2].
The methods are described below in detail.
n+1
(1) Elimination of 0 with first order approximation -
io
single FOE at interface.
The POE for heat conduction in the transformed coordinate
is

ot - a, 0 X + a 2 0 XX (3.7)

From Taylor series, the second derivative is estimated as

(3.20)

where the nodes are shown on Fig. 21. By replacing 0xx


C
by Eg. (3.20) and using an implicit approximation for the time
derivative, we get

n+1 n+1 n+1


be 0 + (1 + be) 0 - 0
n .] (3.21)
C iC-1 ic iC

where

1
ae - ----..",.--...... and
2 a2At
a, At + AX
100
and the coefficients a, and a z result from the coordinate
transformation of the C-side of the interface.
Following a similar procedure for D-side of the
interface, yields

n+1 n+1 n+1 n


Ox = ad [ (1 + bd ) 0 - bd 0 - 0 (3.22 )
D iD i0+1 iD
1
where

1 and ll.t
ad - bd - 2az
2 azll.t (ll.X) z
a,ll.t -
ll.X
and, the coefficients a, and a z are from the
of course,
coordinate transformation on the D-side of the interface.
Equations (3.21) and (3.22) are substituted into the
firf~t interface condition, Eg. (3.13); this results in

+ko
- a b 0 n+1 - -
Xrf d d iD+1
I
xl C ic
koa 0 n
kca 0 n - -
X; d iD
I I (3.23 )

Note that there are more than three unknowns on the left hand
side of Eg. (3.23), so that the resulting matrix is not
tridiagonal.
101

(a) The unknown 0


n+1
iD
is
-
eliminated by a first

order accurate backward approximation of oxic in Eq. (3.14):


this gives

(3.24)

where de is determined from Eq. 3. 18 • Then Eq. (3.23)


becomes

+ -
ko
leo'
ad bd 0
I
n+1
iD+1

(3.25)

(b) 0/ ~~1 can also be determined by forward

approximation of oxlD in Eq. (3.15): this gives


_ 1 dp
oiD - 1 + do 0 ie + 1 + do 0 iO+1 (3.26)
102
where ~ is given by Eg. (3.19). Then Eg. (3.23) becomes

(3.27)

n+l
(2) Elimination of 8 with second order equation - two
iO
equations at interface.
The fini te difference discretization scheme for this
central difference approach makes use of the imaginary
temperatures shown in Fig. 21. The original FOE (e.g., see
Eg. (3.8» applies at point i = iC; therefore

n+1 n+1 n+1 n


b, 8 + b2 8 + b3 8 = 8 (3.28)
iC-1 ic iC+1 ic

n+1
where 8 is imaginary. Using the central difference
iC+1
approximation of 0 [(~x)2] for 8x
c
l
the interface condition

equation can be written as

_ .... _ 8iC+' - 8iC-' ( )


."C 2 ~x - hco 8 iC - 8 iO
103
from which

(3.29)

where de is now defined as

After combining Eqs. (3.28) and (3.29), we get

(b 1 + b3)
n+ 1
I (
0 iC-1 + b 2 -
b ) I
n+ 1 b3 n+ 1 _ nI
d: 0 iC + de 0 iD - 0 iC
I (3.30)

At point i = iD the FDE is

n+1 n+1 n+1 n


b1 0 + b2 0 + b3 0 = 0 (3.31)
iD-1 iD iD+1 iD

n+1
where 0 is imaginary. Using second order approximation
iD-1

(3.32)

with dD defined as
104
Then by cOmbining Eqs. (3.31) and (3.32), the FDE for the D
side is

= 0 I~D
b1 In+1 b ) In+1 In+1
- 0 ie + [ - ~ + b 2 0 iD + (b1 + b3 ) 0 i0+1 (3.33)
do
In contrast to the first order formulations, the tridiagonal
matrix for solution of the interface problem by this
difference formulation has Eqs. (3.30) and (3.33) from which
oiC and 0 iD are determined simultaneously.

3.2.2 Results and Discussion


Both of the first order (backward, forward differencing)
and second order (central difference) numerical techniques
were evaluated by studying the effects of interface heat
transfer coefficient h~, time step At and mesh spacing AX (=
x' AX) on the interface temperatures 0 ie' 8iO and interface
heat fluxes ~, ~ and ~o obtained from computer runs.
In order to evaluate the finite difference techniques as
well as mesh refinement near the interface, the problem
studied was that shown in Fig. 17b with 80 = 1.0, 81 = 1.0,
82 = O. 0 , a = O. 5 and L = 1. 0 • Temperature profiles with
ae = ke = 0.1, aD = ko = 1.0, At· = 0.01 and a refined mesh
having AX = 0.052 at the interface were computed by backward,
forward and central finite difference methods. Results for
hCD = 0.001, 0.1 and 10.0 are shown in Figs. 22, 23 and 24,
respectively.
105

(a) i ;
IIJ. :
I ,

"'~i----t·---+--t- .........-t--+--+---+----( :
I.I~
,
11/ •. I~
,~ I
,::11.'"1
'to ,
: It I .• J
,~ I
,11/I.'"
;tl. : ,

~
,I: ...,
".,
i ~
II
, • JU,
I"
, ,

:~ j;;~
I

1.llj.I--TI--rl-~I~-~I--t;~~;;~~;:~~~~~
•. 0. .... '.11 '.,. •.• V '.Iv .;. ,_ t. (I," •••• •

DISTANCE

(b)

I
,.I~
;,III I
;.IOj
~~ I
,::1 •. '~
;t- I
,a: I.I~
;.,~ I
,11/1.1"
ltl.
,I:
I
•• .,
:111 I
:t- "~
! •. I~
. ..~-~~~~-~-~~~~~~;;~~;:~~~~~
1
I
"'1 '.N .... '.11 ,.U .... ':$0
DISTANCE
t,1O

Fig. 22 Unsteady state temperature profiles for the Type II


interface prototype problem with h~ = 0.001 by
(a) backward and central, and (b) forward difference
methods at the interface.
106

(a)

I
:111
I~
.: :I
II-
;,
;~
,III
;CL
,t •
!III :
:1-
,
tI~
1
; t.f1
"I~': __.-_...-_-.-_--.-_J~~~~~~!!!!~~~~
i, .,.J0.0' .... 0.10 t.U '.7. t." ....
• ,0 0.14 ....
DISTAItCE

(b)

I"
.111
:~
i:l
,'I-
,it
:~
,III
!CL
.,t.
1111
,I-

t." '.'0
DISTANCE

Fig. 23 Unsteady state temperature profiles for the Type II


interface prototype problem with hro = 0.1 by
(a) backward and central, (b) forward difference
methods at the interface.
107

II
(a) I
I

I
!.~J
:Il:I
II-
,::I "'J
;0:: •.•
ill:
:W •.I
J
I'e.J: •• J
llil I

it- "~

i "'L---.--.--~~~
',Ii
0.'
i
I
0,00 '.u '.at ,.U •.• 0 • 10 '.Ct 0,. .... '.to
DIS TAtlCE

(b) 0.1 .
~~~==~~~~~ __ +-~~~__ -+____~____~.

I

t·,L--,----.-~~~
t .•
'.tt 0.11 ,.U •. 10 ' ..0 '.10 'd' I." '.10 1.'0
DISTAHCE

Fig. 24 Unsteady state temperature profiles for the Type II


interface prototype problem with hro = 10.0 by (a)
backward and central and (b) forward difference
methods at the interface.
108

For a very low value of hco (0.001) the forward difference


method (i.e., Eg. (3.27» produced temperatures greater than
80 (= 1.0) for x < a that are obviously erroneous (Fig. 22b).
The other two methods gave correct results (Fig. 22a). with
hCD = 0.1 the problem associated with the forward difference
method persisted (Fig. 23b) for small times, and then with
hCD = 10 it disappeared (Fig. 24b). By examining Eg. (3.27),
it was seen that for very large ~ (when hco is small), the
n+1
coefficient for the 8 term becomes very small and
ic
causes 8iC > 80 (= 1.0). This is because at the interface
the terms a e , be and b d > 0, but ad < o. On the other hand,
n+1
the coefficient for the 8 term for backward difference
ic
(Eg. 3.25) becomes larger with diminishing hw and as a result
no problems are encountered.
Fluxes at the interface are shown in Figs. 25a and 25b.
The flux calculated for the C-side of the interface is denoted
as ge' across the gap at the interface as ~D' and on the
D-side of the interface as CJo. As previously indicated by the
temperature profiles shown in Figs. 22b and 23b, the fluxes
calculated with the forward difference method are negative and
obviously in error for hco < 0.1.For greater values of h co '
all three techniques give comparable results. However, it
appears that either the "backward" and "central" methods give
109

0.2 ,----------------;~~~r---I

0.1 (a)
tI"
atTau-1.0 (with dTau - 0.01 )
"0
CI)
"S - 0 - Backward CIo. qo. Cleo
c- 0.0
E M Central qe. qo. Cleo
0
0

-0.1
--.- Forward qe
- -1Il- Forward qo. qco

-0.2
-4 -3 -2 -1 0 1 2
I0910hCD
0.2

qet qo and qeo by (b)

tI" ~ Backward Difference

-
"0
CI)

::s 0.1
Co
E
--)t- Central and Forward Difference

0
(J

0.0 .l----.--~=:::;:===~---.--_y_----r--r-___r-_r_---.-~
-4 -3 -2 -1 o 1 2

Fig. 25 Heat fluxes at the interface for the Type II


interface prototype problem using various finite
difference schemes: (a) at time = 1.0; (b) at
time = 2.0.
110
methods give the best results because each shows ~ = CJo = ~.

For a longer time, Fig. 25b, all three methods give results
that are almost equivalent for all values of hro'
The strong effect of the interface heat transfer
coefficient hro on the computed interface temperatures is seen
in Fig. 26. As hro increases from 10-3 to. 10 the difference in
the interface temperatures decreases from 1.0 to 0.1.
~rroneous values of 0iC(>1.0) obtained by the forward
difference approximation at the shorter time (t = 1. 0) is
again to be noted. Errors in 0 iC calculated by the forward
difference method are acceptable, however, at the longer time
(t = 2.0).
It was found (Fig. 27) that by increasing 8t from 10-3 to
10-' had a negligible effect on the calculated interface
temperatures. The worst case was for 0 iC at t = 1, which
increased from 0.263 to 0.275 with the rise in 8t. Therefore,
in the interest of computing speed, 8t could be increased
hundredfold without compromising accuracy for the transformed
POE for heat conduction. However, the advantage ceases when
the coefficient a, in Eq. (3.7) is relatively large, and with
a large 8t the stability of the implicit FO procedure is
compromised.
111

0--- ___ ... _ t .. 1.0


... Backward & Centrel Diff•
-- .. - Forward Ditt.

Backward & Central Ditt.


Forward Ditt.

All Dilf. Methods


0.2
---0- t = 1.0
OlD
II t ... 2.0
0.0
-4 -3 -2 -1 0 1 2
log h
10 CD

Fig. 26 Effect of h~ on the computed interface temperatures


for the Type II interface prototype problem using
various finite difference schemes.
112

0.3.,----------------------
•-
......... .,.-*---------~
Sf .
8=
III III .
aiC at t .. 1.0
0.2
~ Backward Diff. . AfJ Difference Methods

~ Central Diff.
CD
- -IJo - Forward Diff.

0.1 o o
0-
o • •
aiD at t .. 1.0 aiD at t .. 2.0
~ AfJ Difference Methods • AfJ Difference Methods

0.0 -t---.------r---..----,----.-----r----.----l
-4 -3 -1 o

Fig. 27 Effect of 8t on interface temperatures for the


Typ6. II interface prototype problem using various
F.D. schemes.
113
3.2.3 Conclusion
Based on the above analysis, the preferred methods for
handling the interfaces appear to be the central and backward
difference methods. It was decided to implement the backward
scheme into the computer program because of its ease of
handling various interfaces. Good results were also obtained
with meshes designed with MRF of 2 to 3 at the interface.
114

CHAPTER 4

MATHEMATICAL MODEL

4.1 The Energy Equation


An axi-symmetric casting is withdrawn from the mold in
the axial direction (x-direction) with a velocity v x • All
convection, relative to the moving casting, is neglected.
Temperature is two-dimensional (i.e., 8(r,x» in the mold and
in the casting. As will be described later, the evolution of
latent heat is taken into account by a modification of the
specific heat in the solidification temperature' range.
Fourier's rate law is assumed, i.e.,

q x - -k ~
ax (4.1)

and

q r - -k ~
ar (4.2)

where qx and qr are the components of the flux in the x- and


r-direction, respectively, and k is the thermal conductivity
(assumed to be isotropic). with all of these assumptions the
energy equation is

(4.3)
115

where p is density and c: is the specific heat.

4.2 Thermophysical Properties and Nonlinearities


Alloys solidify over a range of temperatures with the
volume fraction of liquid decreasing from unity at the
solidification front to zero at the temperature corresponding
to complete solidification. Therefore, in some portion of the
solidifying casting there is a zone of two phases (solid plus
liquid) called the "mushy zone." In the mushy zone the
thermal properties depend upon the fractions of solid and
liquid at a given temperature.
First consider the specific heat of a solidifying alloy.
At a given temperature in the mushy zone the enthalpy of the
alloy, H, is

(4.4)

where
Hs,HL = the intensive enthalpy of the solid and liquid,
respectively
and
fs' fL = the weight fraction of solid and liquid,
respectively.
Because
116

it can be shown that

(4.5)

where
c'p = the effective specific heat of the two phase
mixture,
cps' c pL = the thermodynamic specific heat of the solid
and liquid, respectively,
and
8Ht = HL - Hs·
The term in the parentheses is simply the l-reighted average of
the specific heats. In the all-solid zone of the casting, c'p
simply reduces to cps' an.d, of course, in the all-liquid zone
it reduces to CpL·

For the simulations herein, the density and thermal


conductivity of the alloys were given as a function of
temperature in the all-solid, liquid plus solid, and the all-
liquid zones. These data are presented in a later chapter.
Equation (4.3) is nonlinear because all its coefficients
are temperature dependent. By means of Kirchoff I s transforma-
tion [70], the temperature Q can be changed to a new variable
~, such that

9
~(x,r) - Jk(u)du.
o
117

In this way, the space derivatives of k can be linearized, so


that Eq. (4.3) can then be written as

1 at/> 1 at/> + (4.6)


aat r ar

where a (= k/ pc:) is the thermal diffusivity. After obtaining


t/> values, the 9 values are obtained by a back transformation.
The advantage of the additional computational effort associ-
ated with this linearization procedure is, however, diminished
by the fact that the first order derivatives still have
non-linear terms, and especially in the two phase region the
coefficients of t/>t and t/>x are strongly non-linear.
A more pragmatic method of dealing with the non-linear
conduction terms in Eq. (4.3) is used. By this method

~ (-k ~) - -k a 92 ak ~ (dk)'(~)'(~)
2 2
_ III -k a9 _ (4.7)
ax ax ax ax ax ax2 d9 ax ax

where the derivatives with primes refer to values in the


previous timestep. Similarly

aar (-k ~~) -k :~ - (~~)'( ~~)'( ~~)


a
(4.8)

By combining Eqs. (4.3), (4.7) and (4.8), the energy equation


becomes

(4.9)
118
4.3 Boundary Conditions
The domain for the temperature field is shown in Fig. 28.
The casting, itself, is zone Zl, and it comprises the all-
liquid, the mushy zone (LTS) , and the all-solid regions.
There are three zones that make up the composite mold; these
are Z2, Z3 and Z4. Parts of the mold are exposed to the
ambient environment at Z5 and Z7, and part is cooled by the
water in zone Z6. Downstream from the mold the surface of the
casting is subjected to cooling to the ambient (Z8), and to
secondary cooling by banks of water sprays at Z9 and Z10.
Within the domain, Type I interfaces are prescribed in
the mold between Z2 and Z3 and between Z3 and Z4. All
interfaces between the casting and the mold are Type II.
At the outer surface of the copper mold, cooling is
effected by cocurrent flow of water. The flow rate is high
enough so that the resulting turbulence is assumed to make
temperature gradient in the water in the r-direction negligi-
ble. The water is heated by the heat transfer from the
surface of Zone 4 to the water in Zone 6. Defining the water
temperature as ~, the energy balance for the water gives

a~ _ -v
w
a~ + [ hL ] (0 _ ~) (4.10)
at ax A Pwcw

where
h = the heat transfer coefficient from the mold
to the water,
x ar =0
v
r
1 II! --------------- x(t)
a = a, l!Jl@[!J)OOO> Type II Interfaces
au =0
Type I Interfaces

Z 1

r
3 --
ax =0
.r4 - -

I I
I
I
x1 X
2 x3
I xI
4 XsI x6
Interfaces

L2.:.:~ "~i~
tIC[[]·::jfu1
Parallel to x - axis Parallel to r - axis

Fig. 28 Geometry and boundary conditions for the HCC process. Note the ........
reentrant corners at P, Q and R. \0
120

Vw = velocity of the water,


L = perimeter of the flow channel for the water,
A = cross-sectional area of the flow channel,
Pw = mass density of the water,
and
Cw = specific heat of the water.
The water has an inlet temperature at x = x2 specified as ¢o'
and initial values of ¢(x) are assumed to be ¢o.
The boundary condition that applies to the outer surface
of zone Z4 of the mold at r = r 4 and x2 S x S x3 is

:~ + ~ (0 - ¢) - 0, (4.11)

where ¢ is the temperature of the mold cooling water. This


boundary condition is a Dirichlet type condition.
Dirichlet conditions are also specified along the
remaining lengths of the outer surface of the mold and surface
of the casting. Specifically, where the outer surface of the
mold (Z5 and Z7) and the casting surface (Z8) are in contact
with the ambient, the boundary conditions are written as

2!..
8r + kh (0 - 0., ) - 0 (4.12 )

where 0., is the ambient temperature. Similarly, the boundary


condi tion that represents heat transfer between the casting
121
and the secondary cooling sprays for x4 S x S X6 (Z9 and Z10)
is

-ao + -h ( 0 - 4>8 ) - 0, (4.13)


ar k

where 4>8 is the temperature of the spray water and h is the


heat transfer coefficient for the spray.

4.4 Governing Equations


The partial differential equations and the boundary
condi tions are combined to obtain governing equations amenable
to a computer solution.

4.4.1 Temperature
The solution field considered is applicable to both
axisymmetric cylindrical coordinates (x,r), as well as to two-
dimensional rectangular coordinates (x,z), with the x-axis in
the direction of casting withdrawal. The same PDES can be
written for both geometries provided that a parameter P is
properly selected.
Following Eq. (4.9), the POE for temperature with r > 0
is given by

(4.14)
122

where the subscripted notations for the derivatives are used


for convenience. Recall that the primes refer to values in
the previous time step, except for ~ which represents the
effective specific heat.
Because of symmetry along r = 0,

and l'Hospital's rule gives

, Or
I 1m - - Orr.
r"O r

Then for r = 0, Eq. (4.14) becomes

(4.15 )

Furthermore, p = 1 for cylindrical coordinates (x,r), and


p = 0 for rectangular coordinates (x, z), in which case z
replaces r in Eqs. (4.14) and (4.15) as well as in subsequent
equations.
By defining

(4.16a)

(4.16b)

(4.16c)
123

for r > 0 (4.16d)

for r - r, - 0 (4.16e)

for r > 0 (4.16f)

for r - r, - 0 (4.16g)

then either Eg. (4.14) or Eg. (4.15) can be written as

(4.17)

When mesh refinement by transformation to a (X,R)


coordinate is used following Egs. (3.5) and (3.6) the above
coefficients a, ••• a 4 are rewritten as

(4.18a)

(4.18b)

(4.18c)

for r > 0

0, for r - r, - 0 (4.18d)
124

, 2
k/ (r) ao, for r > 0 (4.18e)

for r - r, - 0 (4.18f)

with a, ••• a 4 defined by Eg. (4.18). The PDE in the transformed


coordinate (X,R) is then written as

(4.19)

4.4.2 Fraction Solid


The effective specific heat, defined by Eg. (4.5), is an
important property in the energy equation for calculating the
temperature field within the mushy zone. Here, the purpose is
to evaluate fL and dfL/dO so that they can be used to evaluate
c~ according to Eg. (4.5).
Because the mushy zone comprises a dendritic solid and an
interdendritic liquid, the microscopic solid-liquid interface
is quite complex. An alloy solidifying in a dendritic manner
is best described by considering a unit volume in the mushy
zone that is small enough so that it has a unique temperature.
On the other hand, the unit volume is large enough so that it
contains both solid and liquid, and the fractions of solid and
liquid in the unit volume can be characterized.
In the absence of convection and diffusion across the
boundaries of the unit volume, good estimates of the fraction
125

solid as a function of temperature can be made. Here two


scenarios are described. In the first, the interdendritic
liquid and the dendritic solid are in equilibrium, and their
relative amounts can be determined from an appropriate
equilibrium phase diagram. In the second, only the solid at
the dendritic interface is in equilibrium with the inter-
dendritic liquid, but because diffusion in the local solid
(i.e., the solid within the unit element, itself) is limited,
the solid behind the interface does not have a uniform
concentration of solute.
Two extremes in the extent of diffusion are considered.
If there is complete diffusion in the local solid, then the
so-called "lever rule" can be used to calculate the relative
amounts of the phases. On the other hand, if there is no
diffusion in the local solid then the Scheil equation can be
applied. Both cases are described in more detail in Flemings
[71], so a derivation of each is not repeated here. The
former case is more realistic when the solute has a very high
diffusivity in the solid (e.g., C, N, B or 0 in steels), and
the latter is more realistic when the solute is a substitu-
tional alloy element and has a very low diffusivity in the
solid (e.g., Mn in steel or Al in an Mg-AI alloy).
126

with complete diffusion in the local dendritic model, the


fraction solid fs is

[0 0
f
s - (1
1
_~) OM -
-
°0 ]
so that

(4.20)

with k p ' the equilibrium partition ratio, defined by

OM is the melting point of the pure solvent, and 0 0 is the


liquidus temperature of the alloy. For solidification with no
diffusion in the solid, the Scheil equation is

so that

(4.21)
127

4.4.3 Interface Coefficients


Following the treatment in §3.2, in order to eliminate
unknown interface temperatures, we define four interface
coefficients. On a Type II C-D interface that is perpen-
dicular to the r-direction (see Fig. 28), the interface
coefficient, d" is

(4.22a)

or after mesh refinement

d, - (4.22b)
heD ~Rr'

where R is the transformed radial coordinate and r' is dr/dR.


The subscript on the thermal conductivity is used to indicate
the thermal conductivity of the material above the Type II
interface. Similarly, d4 is defined as

(4.23a)

after mesh refinement

leo (4.23b)
d4 - ----
hco ~Rr'

where ko is the thermal conductivity of the material below the


interface.
128
On a Type II interface A-B that is perpendicular to the
x-direction, the coefficients are

kA (4.24a)
d2 -
hABAX

with mesh refinement

kA
d2 - , (4.24b)
hABAXx

and

ka (4.25a)
d3 - hABAx

with mesh refinement

d3 -
ks , (4.25b)
hABAXx

where materials A and B are on the upstream and downstream


side of the interface, respectively.
The coefficients d" d2 , d3 and d4 are used in the FOEs
that are described in the next section.

4.5 Finite Difference Equations


Depending upon the location of a node in the temperature
field, the finite difference equation takes on a different
form. As examples, the finite difference equation for
interior points differs from finite difference equations for
129
points along interfaces and at corners. In this section, each
type of node encountered in the entire field is given, along
with-its finite difference equation that is applicable to a
solution of the temperature field by the Alternating Direction
Implicit (ADI) method.

4.5.1 X-Implicit
The nodes at which the unknown temperatures at the next
half time step are to be computed lie on the X-lines of
constant R. These various types of nodes are indicated on
Fig. 29a as Xl, X2, X3, etc., which correspond to the items
numbered as 1, 2, 3, etc. in the following list.
1. Along the centerline, i.e., R = o.
2. At interior nodes.
3. Along a Type-I interface that is parallel to the
X-direction.
4. Along a Type-II interface that is parallel to the
X-direction.
5. Along an exterior interface with heat transfer to
the ambient or water spray with constant and
uniform temperature.
6. At a type I interface that is perpendicular to the
X-direction.
130

Known X1 (Symmetry B.C.)


Inlet --.
Temp. X2 Constant Gradient B.C-+
X4 I'X10
XS
X9--'

(a)

R1

R7 Jl'R10
R9
R1 RS

L R13

(b)
Fig. 29. various types of nodes in the Finite Difference
Mesh for the HCC problem: (a) X-implicit
equations, (b) R-implicit equations.
131
7. At a Type II interface that is perpendicular to the
X-direction.
8. At nodes just downstream of the nodes defined in 7
above.
9. At an exterior interface perpendicular to
X-direction with heat transfer to the ambient.
10,11,12. At reentrant corners R, P, and Q,
respectively.
13. Along the flow path of the cooling water in the
mold.

4.5.2 R-Implicit
In the Alternating Direction Implicit method, the nodes
at which the unknown temperatures are solved lie on the
R-lines (constant X). These types of nodes are shown in
Fig. 29b as R1, R2, R3, etc. and correspond to items 1, 2, 3,
etc. in the following list.
1. At interior nodes (including R = 0).
2. Along a Type I interface parallel to the
R-direction.
3. At a node immediately below a Type II interface and
on a Type I interface parallel to R-axis.
4. Along a Type II interface parallel to the
R-direction.
132
5. At an exterior interface that is parallel to the
R-direction with heat transfer to the ambient.
6. On a Type I interface that is perpendicular to the
R-direction.
7. On a Type II interface that is perpendicular to the
R-direction.
8. At nodes immediately below the Type II interface of
item 7.
9. At the nodes of the exterior interfaces perpen-
dicular to R-axis with heat transfer to the ambient
or water spray with constant and uniform
temperature.
10,11,12. At reentrant corner points R, P, and Q
respectively.
13. Along the flow path of the cooling water in the
mold.

4.5.3 Equations for the X-Implicit Time step (n ~ n + ~)

In the AD! method, the X-derivatives are approximated at


the n+~ time steps, and the resulting temperatures at a
constant R are calculated implicitly based on the temperatures
at constant X at time n. Equations for the various types of
nodes in §4.5.1 are presented in this section. The item
numbers correspond to the items listed in §4.5.1.
133

1. The finite difference equation for temperature at a node i


on the centerline (r = 0) is

n+~ n+~ n+~ n


b1 8 + b2 8 + b3 8 = bs 8
i-1,j i,j i+1,j i,j

n
+ (b4 + b 6 ) 8 ( 4 .26)
i,j+1
The double subscript notation (i,j) gives the nodal coordinate
(x,r). Equations (4.26) to (4.68) are written for the (x,r)
coordinate system of the nonuniform mesh but are equally
applicable for the (X,R) transformed coordinate system of the
uniform mesh.
If central difference is used for the advection term in
Eq. (4.17) or (4.19)

az 2 2az a1 a2 (4. 27a)


--...", b 2 - -t + , b3 - - 2 AX
(AX) 2 A (AX) 2 ~ (AX) 2

If backward difference is used for the advection term

a1 a2 2 a1 2a2 a2
b1 -
AX
, b2 - At AX
+ b3 - - (4.27b)
(AX)2 (AX) 2 ' (AX) 2

But for both cases b4, b s and b6 are given by

a3 a4 2 2a4 a3 a4
b 4 - - 2Ar + b s - At , b6 -
2Ar
+ (4.28)
(Ar)2 ' (Ar)2 (Ar)2
134
The a 1 ••• a 4 in equations 4.27a,b and 4.28 above are in
turn calculated from Eq. 4.16 or 4.18 depending on whether
mesh refinement is used. Note that since at the centerline r
or R = 0, we use Eq. (4.16e,g) or Eq. (4.18d,f) to get a 3 and

2. Interior (r > r 1 = 0)

n+~ n+~ n+~ n n


b1 0 + b2 0 + b3 0 = b4 0 + bs 0
i-1,j i,j i+1,j i,j-1 i,j

n
+ b6 0 (4.29)
i,j+1

coefficients b l above are calculated using Eqs. (4.27a,b) and


(4.28) while using Eqs. (4.16d,f) or (4.16c,e) to define a 3
and a 4 •

3. On X-Line along Type I Interface


135

where for Material C above the interface

fO -

2 2a 2 ] f (4.31a)
[-l\t + (l\X) 2 c 0'

f4 - [- 2a4 ] fo, fs - [- 2.. + 2a4 ] fo


(l\r) 2 c l\t (l1r) 2 c

and for Material D below the interface

go - -f--1--)-,
a3 + 2a4
g, - (2:~ - (l1X)a2 2] 0
go,
l\r 0

(4.31b)
136
4. On X Line along Type II Interface with both x and
r Temperature Gradient Below

(4.32)

where ff, gi are defined by Eqs. 4.31a,b and the second

I
n+~
interface temperature 0 is calculated as follows
i, jD

I
n+~ In+~ In+~
O. •
~,)D
- - d1 0 •
i, )C-1
+ (1 + d 1) o. .
~,)C
(4.33)

with d 1 defined by Eq. (4.22a) or (4.22b). Equation (4.33) is


identical to Eq. (3.24) except that de has been replaced by d 1 •
137
5. On X-Line along Type II Interface with Uniform
Temperature Below

n+% In+% In+~


ke f 1 fJ Ii-1,jC + (ke f 2 + h co ) fJ i,jC + ke f 3 fJ i+1,jC
(4.34)

6. On Type I Interface across X-Line

(4.35)

- [{kAbO[- 2:'r + (A~)2]J. {-~co [- 2:'r


+ + (::)2 JJJ':A,j-l
n
+ r{k
[
Ab kAbo -=.2a
3 -
(Ar)2
4
}
A
- {ke C3 _ keco 2a4 }] fJ/
(Ar)2 B iA,j

+ [{kAbO
138
where

for Material A (upstream of the interface):

(4036a)

for Material B (downstream of the interface):

(4036b)
2
c3 - COo
tot
139
7. On Type II Interface across X-Line with both x
and r Temperature Gradient Downstream

(4.37)

I iB+1,j
n+~
+ c 0
ka z

2a4 1 In
(l1r)2 B 0 iB,j

where b f , c f are defined by Eq. (4.36) above and the second

I
n+~
interface temperature 0 'B ' is calculated as follows.
l. ,)

I
n+~ In+~ In+~ (4.38)
0, , - -dz 0 , , + (1 + d z) 0, ,
l.B,) l.A-1,) l.A,)

with d z defined by Eq. (4.24a or b).


140

8. Immediately Downstream of Point Defined in 7,


along X-Line

b1 oIn+~ + dz + b )
2
oln+~ + b3 oln+~
1 + d2 iA,j 1 + d2 iB+1,j iB+2,j
(4.39)
- b4 oln + bs oln + b6 oln
iB+1,j-1 iB+1,j iB+1,j+1

n+~
In the above equation 0 was replaced by Eg. (4.38).
iB,j

9. On X-Line Perpendicular to Type II Interface with


Uniform Temperature Downstream

- kAb, ~ 2;r + (A~)2] Ol~,j_l (4.40)

+ [kA b3- kAb, L::;2 ]] OI:A,j + kAb, [2:"r + (A~)2] Ol~,j+l


141
10. At the Reentrant Corner Point R

[ kef, - kog, (1 + d,) ] o/n+%


9. . + [ kef2 - kog, (1 + d,) ] 9/n+%.
l.-l,JC i,JC

n+%
+ ke f 3 9 • 1 ·C
/ l.+ ,J

(4.41)

where f i , gi are defined by Egs. 4.31a,b.

11. At the Reentrant Corner Point P

n+% [ h ] /n+% /n+%


-b, 9/ iC-l,jC + b 2 - 2c, + (1 + 1') ; : 9 iC,jC + 2C2 9 iC+l,jC

+ (-2C4 + ..1-) 9/~.


~r l.C,JC+l
+ (1 + 1') ~CD
.~
.
9/~l.D,JD (4.41)

where coefficients b o••• b3 and co ••• ~ are given by


Egs. (4.36a) and (4.36b), respectively, also,

(4.42)
142
Coefficients b i and c i are calculated for the solidifying
billet and ~ = ax/ar in the refined non-uniform mesh. The
n+~
insert temperature 0 is calculated from
iO,jO

In+~

I n+~ ol~o, jO+1 - n


oI iO,jO 0 iO-1,jO
( 1 + ~) ] o.l. 0 , J. 0 - ~ ar +
ax

hco In+~ (4.43)


+ leo (1 + ~) 0 iC, jC

12. At the Reentrant Corner Point "Q"


n+~
-2kob 1 0 •
I • + [2ksbz - leoc1 + hco (1 + ~) ]
l.0-1,)0
o.In+~ .
l.O,JO

n+~
+ kp c2 (1 + d z) 0 •
I l.C+1,)C•
- leo C z d z 0 n+~

I •
l.C+1, )C-1
+ [- 2 ka b 4 + ~ k]
Tr
A 0In
iA, jA-1

-I- 2~leo]
[ - leo c4 + -ar
- In
o. + heD (1 + ~) In .
o. (4.44)
l.B,jB+1 l.C,JC
143

the coefficients b f and c f are calculated for the mold.

The temperature in the solidifying billet at point ic,jC


n+%
i.e. 0 is calculated from
iC,jc

[1 + hco (~+ 1) (,,~ + A'rl] '1::~jC (4.45)

-'I::~ jD + hCD (~+ [,1:::1, jC /


1) AX - 1 'I::~ jC-1 / 1
Ar

13. For Mold Cooling Water

(4.46)
where

h' - (4.47)

4.5.4 Equations for R-Implicit Time step (n + %~ n + 1)

1. Point on R-Line in Interior

b4 ol~+~
1.,)-1
+ ba ol~+~
1.,)
+ b
6
oln+1
i,j+1
(4.48)

- b, O.
I
n+% In+% In+%
• + b 7 O. • + b 3 O.
1.-1,) 1.,)

1.+1,)
144
where

2 2a2 2 (4.49)
b 7 - - -:::-t + , b s - - A"l=
... (AX)2 Ln~

and b, ••• b6 defined by Eg. 4.27a,b and 4.28.

2. Point on R-Line along Type I Interface

(4.50)

where, for the A side (upstream of the interface):

So -
1 s, - [2Ar
a3 a, ] so, s2 - [;t +
2a, ] so'
2a ' (Ar)2 (Ar)2
a, + _2
AX
(4.51a)

s3 - - 3-
[a a,] So ' s4 - [- (AX)2
2a2
] so' s5 - [- 2
- + 2a 2
1So
2Ar (Ar)2 At (AX) 2
145
and for the B side (downstream of the interface):

+ 2a4 ] wo,
wo -
(Ar) 2 ,
(4.51b)

3. On R-Line along Type I Interface just below Type II


Interface (Perpendicular to R-Line)

(4.52)

I iA-1, jD+1
n+~
- -k s 0
A 4

I i,jD+1 + ks
+ -k s
n+~ In+~
( A 5
+
ks w5) 0 W
4
0
iB+1,jD+1
146
n+1
where 0 • •
I
1., JD
is calculated by

oln+1 _ 1 o/n+1 + d4 In+1 (4.53)


i,jD 1 + d4 i,jC 1 + d 4 0 i,jD+1

and d 4 defined by Eg. 4.23a or b.

4. R-Line along Type II Interface with Temperature


Gradient Downstream

(4.54)

/n+~
-
kg
wdO
3 Z I
n+1
iA -1 , j +1
-ksO
A 4 iA -1 , j

n+1
the downstream temperature 0 is then calculated by
iB, j

(4.55)
147
5. Point on R-Line along Type II Interface with Uniform
Temperature Downstream

(4.56)

for j = mr[3] + 1 i.e. at a node just below the mold-metal

n+1
gap replace 0 = 0 In+1
, 'D
l.,J
by
I i,j-1

o/n+1 _ 1 o/n+1 + d4 o/n+1 (4.57)


i,jD 1 + d4 jC 1 + d4 i,jD+1

6, On Type I Interface across R-Line

n+1 /n+1 In+1


- P3ke 0,
/ l.,J, C- 1 + (pz ke - qz ko) 0, , +
l.,J
'b ko 0, , D 1
l.,J +

1
a, az a, az n+~
- Po kc - - - + - qoko - - + 0
[ [ 2AX (AX) Z ] c [ 2AX (AX) z ]D] / i-1, j
(4.58)

+ [POke [~
2AX
+ _az ] -
(AX) 2 c
~ko .
[_a, +
2AX
az ]]
(AX) 2 D
o/n+~
i+1, j
148
where for the C side
(4.59a)

Po -
1
2a4 I
2
P1 - -6t Po IPZ -
[-6t2 + (6r)
2a,]Z
POIP3 -
2a4
POI
(6r)Z
a3 + -
6r

and for the D side


(4.59b)

go -
1
2a4
a3 - 6r
-
I
2
g1 - -6t ~,gz - [2-
6t + (6r)Z
2a, ] ~,Cb -
2a4
(6r)Z
go,

7. On Type II Interface across R-Line with Temperature


Gradient Below

n+1 [
( - P3 k C + gz ko d 1) () "
I i,JC-1
+ pz kc - gz (1 + d 1) ko ] ()"
In+1
"
1.,JC

n+1
(4.60 )
+ Cbko ()"
I 1.,)"D + 1

[ P1 kc + Po kc [ - (6X) c
lln+~
2az Z ] ()"" C
1.,)

+ Poke [ - a1 + az] () In +% - 1 + a z ] () In +
- goko [- a ~
26X (6X)2 C i+1,jC 26X (6X)z 0 i-1,jD

- g1 ko - goko 2a2 ] jln+%


() - ~ko [ - a1 + a z ] () In+%
[ [ (6X)z 0 i,jD 26X (6X)2 0 i+1,jD
149

The new D side temperature 9


n+1
i,jD I
is then calculated
by Eg. (4.33) rewritten for the n+1th timestep.

8. Immediately Below the Point Defined in 7

- b4 9In+1 + [ b -b d 4 ) 9In+1 -b 9In+1


1 + d4 i , j C 8 4 1 + d4 i , j D+ 1 6 i , j D+ 2
(4.61)

- -b 9 n+%
1 I
i-1, jD+1
+ b 9In+%
7 i, jD+1
- b 9In+%
3 i+1, jD+1

n+1
obtained by replacing 9 • •
by Eg. (4.53).
I
~,JD
in Eg. (4.48) for j = jD+1

9. On Type II Interface across R-Line with no Temperature


Gradient Below

9/~+~
n 1
-kCP3 + (kCP2 + h eD ) 9/ +
~,JC-1 i,jC

- kcPo [- 2~1x +
Ll
a2
(llX)
2l9/~+%
~-l,JC
.
(4.62 )
150
10. At Reentrant Corner R

n+1 [
[-P3lee + 'I2ko d,] 0 •
I i,)C-1
+ Pzlee - 'I2ko (1 + d,) ] In+1
0 i .
,)C

n+1
+ Cbko () I i,jC+1

1
a, az n+~
[
- Poke - - +
2AX (AX) z ]1() i-1, jc (4.63)

a, n+~ a, 4az n+~


-~ko -Tx [ 1
)1()i-1,jD- [ [Tx+
qoko (AX)z ] +q,ko ]11
()i,jD

11. At Reentrant Corner P

n+1 [ heD ]In+1 In+1


-2P3 () I iC,jC-1 + 2pz - qz + lee (1 +~) () iC,jC + Cb () iC,jC+1

- (-2 P 4 +
AX
2...) ()I~+~ . + [2(P1 + P4) + (-q1 +~) _
l.C-1,)C
3~] ()In+~
AX iC,jC

+ ( _~ + 2~ ) () n+~
I (1 + ~) () In+~
h
+ ~
Ax iC+1,jC lee iD,jD (4.64)
151

where
>. _ Ar
AX

in the refined non-uniform mesh. To obtain the insert


n+1
temperature 0
IiD,jD
at the reentrant corner P we use:

n+1 In+1
n+1 0I iD,jD+1 +' 0 iD-1,jD
2:.. + (1 + >.) O. • - llr AX
[ llr ] I ~D,JD
A

+ hco (1 + >.) 0
In+1
(4.65)
ko iC,jC

12. At Reentrant Corner Q

n+1
+ 2ko~ 0I iD, jD+1

- + 2>.ka] 0 In+~
[- k
AP4 --xx iB-1, jB
(4.66 )

+ [ - 2ko ~D + ->.ko)
AX
I
o.n+~
~D+1,JD
. + hco (1 + In+~ •
>.) O.
~C,JC
152

coefficients Pf are calculated for the break-ring and qf for


the mold. The temperature of the solidifying billet 01~+l,
~C,)C
at point Q is given by

1:.. + hco (1 + A) ] o/n+1, _ 1 oln+1 + hco (1 + A) oln+1


[ llr ko iC,)C llr iC, jC-1 kc iD, jD

,\ [ In+% In+%] (4.67 )


llX 0 iC+1,jC - 0 iC,jC

13. For Mold cooling water

vw
- EX ¢
In+1
i -1 +
(2II t +h
,) ¢ In+1
i
vw In+1
+ 211x ¢ i +1

2 In+% In+1
- - ¢ + h' 0 (4.68 )
llt i i

where h' is defined by Eq. (4.47).


153

CHAPTER 5

PROGRAM STRUCTURE AND VALIDATION

5.1 Program Structure


5.1.1 Capabilities
The mathematical model described in Chapter 4 was
programmed into a simulation package that was designed to
treat variations in billet shape (round or slab), three
different alloys (Mg, Al and Fe alloys), various mold
materials including a break-ring, different mold designs
(straight or stepped molds), and operating conditions (e.g.,
steady or cyclic withdrawal). The package is named CASTSIM-PC
and consists of over 20,000 lines of code in PASCAL.
Five types of continuous casting molds and cooling
arrangements can be handled by the code by setting the
variable MoldType between 1 and 5 (Fig. 30). Uniform or
non-uniform refined grids can be chosen by setting
coefficients in the grid refinement equations.

5.1.2 Organization of Modules and Programs


The simulation code is separated into three major parts,
namely Preprocessing, Runtime and postprocessing modules. The
modules are shown in Fig. 31 in the order that they are used
to solve a problem. All data entry is via windows which check
154

Z1
MoldType=5
Z2

Z5 Z6

Zt MoldType=4

Z8
Z4 7:T
Z5 Z6 Z8 Z9

Z1 MoldType=3
f:

Z1
MoldType=2

Z1 MoldType=1

1----- 26--, zo
Z8 Z9

Fig. 30 Mold design capability of the continuous casting


software CASTSIM-PC.
155

entries for type of variable (integer, real, boolean) and


range as well. A data base (Therm2) containing coefficients
of temperature dependent thermophysical properties for up to
20 materials (casting alloys, mold, insert, water) along with
the solidification properties of the casting alloys is
included as part of PreB4 in the preprocessing module.
For problems that have already been set up, minor changes
in operating conditions and simulation parameters can be made
from the runtime module RunB4. If it is so desired, results
from a previous simulation run can be used as initial values.
RunB4 can be executed in a debug mode for a limited number of
time steps, when values from consecutive half (x-implicit) and
full (r-implicit) time steps are saved to aid in checking
results.
The temperature and fraction solid are plotted in three
dimensions by the postprocessing program called PostB4. The
plots are helpful in identifying trouble spots, such as
growing instabilities and their effect on neighboring field
values. This program also plots specified isotherms: if none
is specified then the liquidus and the solidus (or eutectic
temperature) in the cast billet are plotted as the default
isotherms. In addition, the positions of the solidus and the
liquidus at 'the billet surface and center are also printed.
The program HistB4 scans the data saved at various times and
156

preProcessing

MshGn
win4
PreB4

RunTime

RunB4

PostProcessing

PostB4
HistB4

Fig. 31 Major components of the continuous casting


simulation software CASTSIM-PC.
157
creates time histories of heat input and output, and
temperatures at specified nodes.

5.1.3 Brief Description of Programs


Various components of the CASTSIM-PC code package are
described below.
PREPROCESSING
MshGn: This program is used to generate the non-uniform
refined mesh for both the x and r axes by the
coordinate transformation method described in
Appendix A. Depending upon the constraints and
input data for the desired mesh, coefficients of the
transformed equations are computed using one of the
following four variations of MshGn.
MshGn1: This version of MshGn is used when the only
constraints governing the mesh are that it
should have maxima/minima in mesh spacing (6X
or 6r) at certain locations. Input data in
this case are the number of maxima/minima,
their locations in both the original and
transformed mesh (x, X or r, R) and coordinate
of the endpoints.
MshGn2: If in addition to the constraints for MshGn1,
we also require that the transformed mesh
should also map certain internal points of the
158

original mesh, then MshGn2 is used. The


additional input data required are the number
and coordinates (e.g., x, X) of the internal
points.
MshGn3: When we need to force the mesh spacing ax (or
ar) at the origin (e.g., x = X = 0) to be a
certain value while still satisfying the
constraints discussed above, then this version
of MshGn is used. Additional input data is x,
at x = X = o.
MshGn4: This program is used if, in addition to all of
the above constraints, we also require a
minimum in mesh spacing at x (or r) =0 which
adds the condition x" = 0 at x = 0 to the

input data.
Win4: This sets up a window on the monitor screen that is
used for the entry and editing of input data per-
taining to billet and mold geometry, billet and mold
materials, coefficients for grid refinement, heat
transfer coefficients, initial and boundary values
of temperature, operating conditions for the casting
operation, and various simulation and graphics
display parameters.
PreB4: This program reads the input data created in Win4
and initializes all parameters. Depending on type
159

of mold (Fig. 30), each node is assigned numbers


that signify material, and the type of finite dif-
ference equations (X-implicit and R-Implicit) for
that node. suitable thermophysical properties are
obtained from the Therm2 data base. This program
also sets initial values of temperature and fraction
solid.
RUNTIME
RunB4: The operating conditions and simulation parameters
can be changed if needed with this program. Then
the simulation is carried 'out. After the interior
temperatures are obtained by solving the tridiagonal
matrix arising from the ADI method used here, the
interface temperatures are calculated by the appro-
priate equation. During execution, a heat balance
at each timestep can be done to check the solution
for accuracy. The option of selecting either a
constant withdrawal or cyclic withdrawal is also
available. The calculated temperatures and fraction
solids after full time steps at required time
intervals are saved for postprocessing.
POSTPROCESSING
PostB4: This program produces graphic displays on the
monitor screen for temperature, isotherms, and
fraction solid that were created and saved by RunB4.
160
Hard copies of plots and numerical data can also be
made.
HistB4: Using data created and saved by RunB4, this program
makes plots of temperature as a function of time.
These can be viewed on the monitor screen, and hard
copies of plots and numerical data can be printed
if required.

5.2 Verification
The simulator code, CASTSIM-PC, was verified for the
continuous casting of Mg-AI alloys [30] and Fe-0.7% C steel
[35] •

5. 2 . 1 continuous C,asting of Mg-AI Alloy


The measured and computed temperatures for the casting of
a 20.32 cm radius billet of Mg-8% Al alloy, as reported by
Adenis et ale [30] and discussed in §2.3.1, were used to
validate CASTSIM-PC. Their paper contains excellent docu-
mentation of input data including thermophysical properties,
some of which are reproduced in Table 7. However, they failed
to report the mold-metal heat transfer coefficient that gave
the best agreement between the measured and calculated
temperatures.
161
5.2.1.1 Estimation of Mold-Metal Gap Heat Transfer
Coefficient
Values of the heat transfer coefficient (haop ) along the
length of the mold were estimated by three different methods:
Method 1: The model proposed by Adenis et al. [30] for
calculating the mold-metal gap thickness from
shrinkage of the billet and for estimating the
gap heat transfer coefficient was used. It is
based on conduction and radiation through the
air and oil in the gap. The thermophysical
properties and measured billet surface
temperatures given in Fig. 15 were used for
these calculations.
Method 2: The radial heat flux from the billet to the
mold was calculated. This flux was based on
the surface and sub-surface temperature
measurements of Adenis et al. [30] (Fig. 15).
They also gave the temperature of the mold
next to the casting so that hoop could be
calculated.
Method 3: In their Fig. 7, Adenis et al. [30] reported
values of the radial heat flux into the mold
cooling water. Then with their measured
billet surface and cooling water temperatures,
an overall heat transfer coefficient was
162
Table 7
Input Data to Simulate Run AB of Adenis et al. [30]

Billet Radius: 20.32 cm


Casting Speed: 0.089 cm. s"1
Alloy: AZ80A (Mg with 2.5% Al, 0.5% Zn, 0.2% Mn)

Melt Temperature: 674°C


Liquidus Temperature: 604.4°C
Nonequilibrium Solidus Temperature*: 437°C
Partition Coefficient kp (Mg-Al Binary): 0.41
.6Ht : 78.9 cal.g"1
p: 1. 775 - 1. 545 X 10"4 (}, g cm"3
cp: 0.246 + 1.0656 x 10"5 8, cal.g"1.o C"1
k: 0.12285 + 0.000234 8, cal. s"1. cm"1. ° C"1
Length of Solution Field: 44 cm
Mold Length: 24 cm
Mold Thickness: 0.635 cm
Mold Material: copper
p: 27 < 8 < 2000°C, p = 8.94 g cm- 3
cp: 27 < 8 < 727°C, c p = 0.0906 + 0.239 x la- 4 () cal.g- 1.cm- 1
k: 27 < 8 < 900°C, k= 0.956 - 0.22 x 10-3 8 cal·s- 1·cm"1.oc- 1
Heat Transfer Coefficients
Metal-Mold Gap: see Fig. 32
Mold-Cooling Water: 0.542 cal·cm"2. s "1.o c -1
Billet-Cooling Water: 0.27 cal. cm- 2. s"1. °c- 1

*Assumed to be equal to the eutectic temperature in the


Mg-Al system.
163

0.07

,.. AIR
0.06

0 I
0
,.. 0.05
I
• (/) I h calculated by
I gap
·E
C'II
I • Method 1
u
B
0.04
I
I
.. Method 2

$l,
0.03 I
I
• Method 3

.c
1\1
C'I I
0.02 I
r-
I
0.01 L_
0.00
0 10 20 30

Distance along Mold, em

Fig. 32 Calculated mold-metal gap heat transfer coefficient


for Run AB from Adenis et ale [30]. Extrapolations
used in calculations are shown as broken lines and
explained in the text.
164

calculated. From this overall heat transfer


coefficient, the conductivity of the mold, and
the heat transfer coefficient between the mold
and the water, a simple calculation for the
heat transfer coefficient (hoap )' applicable to
the casting-mold interface, was done.
The values of hoap along the mold obtained by these three
methods are plotted in Fig. 32.
A limitation of the gap thickness model of Adenis et ale
[30] is that when there is no shrinkage gap between the metal
and mold, hoap becomes infinitely large. This trend is seen in
Fig. 32 for hoap obtained by Method 1. In this case, it is
assumed that hoop is uniform at the entry length ex < 3.3 cm)
as recommended by Michalek and Dantzig [35].
The values of hoap calculated by Method 2, on the other
hand, do not increase rapidly at the entry. There is an
apparent increase near the mold exit, as the influence of the
direct-chill water contacting the billet below the mold is
fel t upstream as well. As a result, near the exit of the mold
the billet surface temperature drops rapidly thereby
invalidating the simple method of calculating the radial flux
based on one-dimensional conduction. Therefore, for the
purpose of simUlation this rise in the calculated hoap near the
mold exit is neglected, and extrapolated values from upstream
are used instead.
165
Near the top of the mold both Methods 2 and 3 show a
slight drop in hgOp. For simulations using h llop by Methods 2
and 3, above the position at which the maxima in the calcu-
lated heat transfer coefficient hoop occur, the h llop values were
taken to be equal to the respective maximum.

5.2.1.2 Simulation Results and Analyses


The simulation of the steady state temperature in the
experimental casting of Adenis et ale [30] is shown in
Fig. 33. Method 3 was used to evaluate hgOp (Fig. 32). Note
the use of grid refinement in both the x- and r-directions.
Casting withdrawal is from right to left, and there is a water
cooled mold for 0 S x S 24 cm. The large discontinuity in
temperature, which appears as the side of a cliff in Fig. 33,
represents the temperature drop between the surface of the
billet and the surface of the mold. The temperature in the
mold is at the bottom of the cliff. At x = 0, the temperature
at the surface of the mold is greater than 500 C, but it drops
G

rapidly to almost the temperature of the cooling water in the


first 5 cm. There is a sharp drop in the surface temperature
of the billet, as it emerges from the mold and contacts the
cooling water.
The approach to steady state is shown in Fig. 34. The
temperature at the centerline of the bottom of the field was
selected because it is the slowest in reaching the steady
···.a ·a •••

MOLD t8
~

Fig. 33 Simulation of steady-state temperature field in the billet and mold


for Run AB of Adenis et ale [30] by the ADI method. Method 3 was
used to calculate h gap • ....0\
0\
167

500~------------------------------------------~

"
(J
o 400

~
:J
!CD
Co
E
{!!. 300

• At Point 1

~ AtPoint2

200+---~--~~--~--~--~--~--~----~--~--~
o 100 200 300 400 500 600
Simulation Tlmesteps

Fig. 34 Approach to steady state in the simulation of the


experimental casting of Adenis et al. [30 ] with
6t = 2s. Temperature at (1) billet center, 44 cm
from inlet end, and (2) billet surface, 25 cm from
inlet end.
168

state. The simulation time required to reach this steady


state (800 s) is somewhat longer than the time taken by metal
to move from the inlet to the exit end of the solution field.
This indicates that, even for this relatively conductive
alloy, heat flow in the withdrawal direction by advection
predominates over longitudinal conduction. Moreover, no
oscillation of temperatures, especially at Point 2
(corresponding to reentrant corner R of Fig. 28), was seen,
indicating a stable numerical scheme.
The approach to steady state is also seen in Fig. 35a
where heat input to and output from the solution field is
plotted as a function of simulation time. The quantities
plotted are:
qlnH: heat in by advection and conduction at the inlet
over the X-implicit time step n ~ n + ~.

qOutH: heat out from the billet surface to mold, to the


ambient and to the secondary cooling water as well
as the heat carried out by the solidified billet by
advection and conduction at the outlet end of the
solution field. Note that qOutH is defined for the
X-implicit time step n ~ n + ~.

qlnF: same as qlnH but for the R-implicit time step


n+~~n+1.
169

-c.
GI
rGIn
E
50

-
i=
i ii
(.)
..:.:

-
::J
c.
'5
40
qlnH
qlnF

--
0

:s
c.
0=-
II
qOutH
qOutF

-
.5
co
GI
:z:
30
0 200 400 600 800 1000 1200
(a)

60

Ii 50

-
.!!!
.5 40
E
-B
~

c
0
30

-co
en
is
20
--a-
o
Solidus at Billet Col
Liquidus at Billet Col

10
0 200 400 600 800 10·00 1200
Simulation Time. s

(b)

Fig. 35 Time history of (a) computed heat input/output and


(b) position of liquidus and non-equilibrium solidus
at billet centerline for simulation of the
experimental casting of Adenis et ale [30].
170
qOutF: same as qOutH but for the R-implicit time step
n+~-'n+l.

From Fig. 35a it is seen that steady state is reached in


600 to 800 s, similar to the results shown in Fig. 34. Note
also that qInH is smaller than qInF by less than 1%, at steady
state qOutF is larger than qInF by 1.5%, and qOutH is smaller
than qInH by 4%. Ideally they should all be equal. at steady
state. The differences can be attributed to an underestimate
of the temperature gradient along the r-axis at the mold-metal
interface after the X-implicit step. These small deviations
can be reduced by jUdicious mesh refinement.
In Fig. 35b the positions of the liquidus (604.4°C) and
the non-equilibrium solidus temperature (i.e., the eutectic
437°C) at the billet centerline are plotted as a function of
simulation time. steady state is reached in 600 to 800 s.
The importance of choosing the correct hgap for heat
transfer from the billet to the mold is clearly seen in
Figs. 36 and 37 where simulated steady state temperatures,
based on the hgap profiles of Fig. 32, are compared to measured
temperatures. For both billet surface (Fig. 36) and sub-
surface temperatures (Fig. 37), hgOp based on Method 3 gives
best agreement with measurements. The sharper drop in billet
surface temperature (Fig. 36) upstream of the end of the mold
when compared to the measured values may, in part, because of
the use of too high a value of billet-water heat transfer
171

700~--------------------------------------------~

600

(,)
o 500

~
-
::s
f!
Q)
Co
-e-- Measured [30]
E 400
{! Computed in this work.
with h by:
gap
• Method 1
300 ~ Method 2
~ Method 3

200+-------~------~------~------~------~---~--~
o 1a 20 30
Distance from Meniscus I cm

Fig. 36 Comparison between measured and computed


temperatures on the billet surface.
172

700~----------------------------------------------'

0 600
0

~
-E
::J

CI)
• Measured (30)

computed in this work


Co
E with h gap by
~
500 ~ Method 1

a Method 2

~ Method3

400+-------~----~r-----~------~------~------~
o 10 20 30

Distance from Meniscus I cm

Fig. 37 Comparison between measured and computed


temperatures 5.08 cm under the billet surface.
173
coefficient (0.27 cal·cm- 2 ·s- 1.oc- 1 as reported in Adenis et ale
[30]).
Liquidus and solidus (eutectic) isotherms in the computed
temperature field were plotted (Fig. 38) and compared to the
experimental measurements of Adenis et al. [30]. The computed
liquidus isotherm matches the measured liquidus profile quite
well with the largest difference at the centerline of the
billet. It is also seen from Fig. 38 that the solidification
is not complete at the exit end of the solution field.
However, this does not imply that errors were introduced into
the simulation by the choice of too short a solution field
(44 cm). This was verified by repeating the simulation with
a longer solution field of 59.3 cm. The difference in
temperature values and the position of liquidus between the
two simulation results were negligible.
For this relatively conductive Mg-AI alloy, secondary
spray cooling in the submold region affects the rate of
solidification even in the billet centerline, and therefore
the correct choice of spray heat transfer coefficient is
important. The above mentioned difference in the measured and
calculated liquidus positions at the billet centerline appears
to be due to the spray heat transfer coefficient (0.27 cal cm- 2
s-1 ·c- 1) quoted by Adenis et al. [30] being too high. When
the simulation was repeated with a spray heat transfer
coefficient of 0.0675 cal cm- 2 s-1 ·c- 1, the position of the
174

x
.....
,=
t:
=:::::
: : : : ::::
, : : :::::. r

·:~:·~::f::~:~::~:::~::t::~::::.::~:::~::~::~:·::::t:··:·::'·f·:~:~::7~ ~~:
~1~~-SECONDARY--~~~I~4----------MOLD------~~
COOLING

Fig. 38 Comparison between the experimental (CUrve 1) and


simulated (Curve 2) liquidus isotherms. Also shown
is the simulated solidus isotherm.
175
calculated liquidus at the billet centerline moved downstream
and agreed with the measured position of Adenis et ale [30].

5.2.1.3 Numerical Results and Analyses


The two parameters critical to efficient running of the
unsteady state simulation are the time step and grid spacing.
Being implicit the ADI method is unconditionally stable (i.e. I
errors remain bounded irrespective of ~t, ~x and ~r) but the
accuracy is still O(~t, (~x)2 and (~r)2). It is desirable to
determine accuracy as a function of ~t, ~x and ~r to minimize
execution time. Heuristic analyses of the error bounds
provide a practical alternative to a rigorous analysis for
these continuous casting problems.
By running the simulation with ~t ranging from 0.1 to
10.0 s for 100 s, the effect of the timestep on the accuracy
of the solution was determined (Fig. 39). Errors in the
solution at a particular mesh point were defined as the
difference between the temperatures obtained with a particular
~t and with ~t = 0.1 (considered the most accurate). The
errors were tracked at points near the mold as well as at the
center of the bottom of the temperature field. Up to ~t = 4.0

errors at all these points increase very slowly and remain


within 1°C: for ~t > 4.0 the errors increase substantially.
Moreover I small oscillations in the temperature profile
176

10~--------------------------------------------~

-------- ------------
~~=_~~~~==:~~WY~~.~ftft~-~~-~~::.~~:~_~------------------- .-_.
o .. _....-... -._ ... _.-._ ......
0 0
....
E
Point 1
w •••••• - Point 2

·10 Point 3
,
I Point 4
'.1

·20~--~--~~--~--~--'----~---r-~--~--~--~
o 2 4 6 8 10 12

Simulation TlmeStep 4 t. s

Fig. 39 Effect of At on the calculated temperature at


selected locations in the billet.
177
appeared on the billet surface. For subsequent runs
At = 2.0 s was used without any problems.
The effect of mesh refinement on the calculation of
radial temperature gradients was mentioned in the discussion
for heat balance in §5.2c1.2. The effect of mesh refinement
in terms of the mesh refinement factor MRF (defined by
Eg. 3.10) on the computed temperature along the surface of the
billet is seen in Fig. 40. Billet surface temperatures
reported by Adenis et ale [30] show a steep drop as the billet
emerges from the mold and is subj ected to secondary spray
cooling. Because of the finite difference method used in this
work, the effect of the spray cooling is transmitted to the
mesh points upstream of the mold exit, and as a result the
computed billet surface temperatures start dropping ahead of
secondary cooling. However, the results plotted in Fig. 39
show that the difference between the computed and measured
billet surface temperature can be greatly reduced by local
mesh refinement at the mold exit.

5.2.2 vee of Fe-0.7% e steel


The test problem chosen was that used by Michalek and
Dantzig [35], who demonstrated a method for calculating the
mold gap thickness and billet-mold heat transfer coefficient
(hoop) in vee of Fe-O. 7% e steeL Input data for this simu-
lation is given in Table 8. simulations were run with their
178

700,--------------------------------------------

600

au --'0---
500
!
-
:::I
f!
CI)
a.
• Measured [30]

computed In this work with


E 400 h by Method 3
CI) gap
I-
~ Uniform Mesh

--0-- Non Uniform Mesh (M R F .. 1.4)


300

200+-----~------~----~~----~------~----~
o 10 20 30
Distance from Meniscus, em

Fig. 40 Effect of mesh refinement on computed temperatures


on billet surface, Adenis Run AB.
179

input data and their values of hgap that are shown in Fig. 10.
Because Michalek and Dantzig [35] did not give a sub-mold heat
transfer coefficient for cooling of the billet by the ambient
or water-sprays, the lowest value of hgap from Fig. 10 (at mold
exit) was used for the sub-mold region. The sub-mold heat
transfer coefficient thus obtained was 0.02 cal. cm- 2 • s-1 . C- 1•
0

As it turns out this is also the value of the heat transfer


coefficient for water spray cooling quoted by Larrecq et ale
(49]. Mesh refinement was used for both the x and r axes.

5.2.2.1 Numerical Difficulties


The two orders of magnitude increase in vx/a for this
test problem compared to the previous case (§5.2.1) of the
Mg-Al alloy is significant because it may lead to numerical
difficulties. When Ox in the advective term of the PDE was
discretized by central difference, gross oscillations appeared
in the solution.
The unconditional stability of the ADI scheme in either
time or space domain is guaranteed only for diagonal dominance
of the resulting tridiagonal matrix (66,72]. Just as this
dominance may be compromised by certain boundary conditions
(73], the use of central difference for the advective term
caused the oscillations observed for this test case. The
oscillations were localized in the two phase region near the
centerline where the conduction terms were small. In
180
Table 8
Input Data to simulate VCC of Fe-0.7%C Steel [35]

Billet Radius: 4.1 cm


casting Speed: 2.116 cm. s-1
Alloy: Fe-0.7%C Steel
Melt Temperature: 1530°C
Liquidus Temperature: 1486°C
Solidus Temperature: 1398°C
27 < 8 < 900°C, p = 7.8 g.cm- 3
900 < 8 < 1300°C, p = 8.17 - 0.6 x 10-3 8
1300 < 8 < 1600°C, p = 10.68 - 0.254 x 10- 2 8
27 < 8 < 1500°C, c p = 0.1286 + 0.2443 x 10- 4 8
cal. g-1 ° C-,
1500 < 8 < 1600°C, c p = 0.165
k: 167 < 8 < 882°C, k = 0.1235 - 0.64 x 10- 4 8
cal s-,cm-1 ° c- 1
882 < 8 < 1522°C, k = 0.0455 + 0.2437 x 10-4 8
1522 < 8 < 1700°C, k = 0.826
in the solidification range i.e. 1398 ~ 8 ~ 1486°C,
the latent heat of melting is incorporated into
c p to give c~ = 0.827 cal.g- 1. o c- 1
Length of Solution Field: 50 cm
Mold Length: 45 cm
Mold Thickness: 2.7 cm
Mold Material: copper (for properties see Table 7)
Heat Transfer Coefficients
Billet-Mold Gap: see Fig. 10
Mold-Cooling water: 1. 0755 cal. cm- 2 • s-1. °c- 1
Mold Air: 0 • 003 cal cm- 2 s-1 ° C- 1
Billet-Water Spray: 0.02 cal. cm- 2 • S-1 . C- 1
0

(1)Equations for k, Cpo and p were obtained from numerical data


given in Table 1 of Ref. 35.
181

addition, the release of latent heat in this region increases


the coefficient of the advective term, and the central
differencing for this term prevents diagonal dominance in the
matrix.
An examination of the finite difference approximations
[Eq. 4.27] shows that adequate reduction of the timestep ~t

can restore diagonal dominance; however, use of non-centered


differencing of the first-order derivative in the advective
term, as suggested by Spalding [74 J and Peyret and Taylor
[73], is more economic. Since Vx > 0 in this case, pure
backward (upwind) differencing was used and the oscillations
disappeared (Fig. 41).

5.2.2.2 Results and Analyses


Michalek and Dantzig [35J presented their results in
terms of the "shell" thickness, defined by the distance from
the mold surface to the solidus isotherm and to the liquidus
isotherm. In the initial stages of shell solidification, it
was necessary to use mesh refinement to keep ~t comparable to
that used by Michalek and Dantzig [35J. The shell thickness
calculated by them and with CASTSIM-PC are given in Table 9.
With a uniform mesh (DZ1), the shell thickness 50 mm below the
meniscus does not compare very well to the results given by
Michalek and Dantzig [35J; however, at the mold exit they do
agree. With mesh refinement (DZ5 and DZ3) there is good
182

~
x r

Fig. 41 Calculated temperature plot in the billet and mold


for VCC of Fe-O.7% C steel. Conditions given in
Table 8. Upwind difference for the advective term
in PDE during X-implicit calculations.
Table 9
simulations of vee of Fe-0.7% e steel

Shell Thickness I

~r 50 mm below meniscus at mold exit I


Interface
Run Mesh mm Liquidus Solidus Liquidus Solidus I
mm mm mm mm i

,
Michalek 51x46 0.66 4.21 1.45 16.0 8.75
& Dantzig
[35]
DZ1 56x22 3.417 5.95 3.177 15.54 8.69
DZ5 56x24 1.11 4.23 1.96 14.68 8.104
DZ3 56x17 1.18 4.028 1.98 15.66
-_.- -----
8.4
1. For DZ1, DZ5 and DZ3, ~t = 0.25 s, t = 30.0 s.
2. DZ1 has uniform mesh spacing with ~x = 10.0 mm, ~r = 3.417 mm.
3. DZ5 and DZ3 have non-uniform meshes in both x and r axes. For both runs,
at 50 mm below the meniscus ~x = 5.5 mm and at mold exit ~x = 15.3 mm. At
the billet-mold interface ~r = 1.11 mm for DZ5 and 1.18 for DZ3.

~
C)
w
184
agreement all along the shell. However the finer mesh (DZ5)
underestimated the shell thickness by a maximum of 8 percent
compared to that calculated by Michalek and Dantzig [35].

5.2.2.3 Conclusion
It was established that to accurately simulate continuous
casting of steels (when vx/a is relatively large), it is
necessary to use backward differencing for the advective term,
and mesh refinement in the radial direction to accurately
compute the shell thicknesses in the initial stages of
solidification.
185

CHAPTER 6

SIMULATION OF HORIZONTAL CONTINUOUS CASTING

6.1 Choice of Computational Meshes for HCC


In addition to Mesh Refinement, discussed in Chapters 3
and 4, further considerations need to be taken into account to
select optimum meshes for the simulation of HCC of steel. In
the derivation of Eq. (4.3), i.e. the advection diffusion PDE
for continuous casting, it was assumed that within any
infinitesimally small volume element in the solution field,
complete equilibration takes place between the material
occupying that element and any new material that flows into or
out of it during a short time interval. The mesh employed for
the finite difference solution of this equation can be either
fixed in space or move with the casting itself. Both have
their advantages and disadvantages.
If a fixed mesh is employed then it is as if the liquid
and the solidified shell of the casting flow past individual
grid points. In the case of well-mixed flows with no solids,
it is easy to see why the assumption of complete equilibration
holds even after Eq. (4.3) is discretized into the finite
difference form using large volume elements/grid spacing. For
both VCC and HCC, after solidification begins Eq (4.3) has to
account for liquid, as well as the growing shell containing
186

it. For flows involving rigid solids with low thermal


conductivity, i.e. the shell in vee and Hee of steel, the
equilibration assumption may fail even if the finite
difference discretization uses small grid spacing or short
time intervals. If non-uniform refined fixed meshes are used
to solve various vee and Hee problems, then in some regions of

the mesh the grid spacing 8X may be large enough so that


8X > Vx 8t. If this happens, then the finite difference
methocl imposes unrealistic mixing in the rigid solid contained
within the length 8X. Such a situation will also arise for
uniform fixed meshes with coarse grid spacing.
The spurious smoothing of a field variable, e.g.,

temperature in the solid shell of the billet, caused by

excessive "numerical mixing" or "numerical diffusion" [75]


introduces errors that can be unacceptable for simulating Hee.
Because of such "numerical mixing", the alternating crests and
troughs in the shell formed by cyclic withdrawal in Hee become

smeared in the finite difference solution obtained with a


fixed mesh of inappropriate grid spacing.

A moving mesh does not introduce "numerical mixing" into


the solution and may therefore appear superior to a fixed mesh
for simUlating solidification of the shell during vee or Hee
of steel billets. The use of a moving mesh will convert a
continuous casting problem into that for a static casting
growing towards the inlet end. The vee or Hee problem can
187

then be solved by setting the withdrawal velocity to zero and


all motion of the billet simulated by movement of the mesh.
For inherently transient problems, e.g. HCC, new grid points
with updated values of field variables may be added at the
inlet end at a rate dependent on the withdrawal velocity of
this billet. Old grid points may be removed from the outlet
end if a constant field length is desired. The addition and
removal of grid points may be simulated by shifting the field
variable values to adjacent grid points along the flow
direction at suitable time intervals.
Simulation of HCC by Holleis et a1. [40] and by Pehlke
and Ho [41], discussed earlier in §2.3.4, used combinations of
fixed and moving meshes. A fixed mesh for the newly forming
shell attached to the break-ring was combined with a moving
mesh for the moving shell downstream of the secondary witness
mark (SWM). This mark also formed the boundary between the
two meshes. At each timestep a number of gridpoints (deter-
mined from distance travelled by the moving shell divided by
grid spacing) were added to the moving grid at this mark.
Temperatures at the new grid points were made the same as that
at the boundary gridpoint located at the SWM and the simula-
tion continued.
A consequence of using this combined mesh by the above
mentioned authors [40,41] is that numerical mixing in the two
phase zone downstream of the SWM is eliminated. As a result
188
the shell profiles calculated by Holleis et al. [40] and by
Pehlke and Ho [41] had pronounced waves. The latter authors
neglected relative motion between the liquid and the solid
shell downstream of the SWM. On the other hand, the experi-
mental shell profile reported by Taylor and 1,alner [26] and
shown in Fig. 5, exhibits a waviness that is less pronounced,
and the shell adjacent to the break-ring showed considerable
remel ting. This indicates that in the actual HCC process
there is fluid flow past the SwM and mixing along the billet
length. In the absence of an explicit calculation of liquid
velocities, "numerical mixing" associated with a fixed mesh
may actually give more representative results, whereas the use
of a moving mesh may be too restrictive in terms of mixing.
In the present simulation of HCC, a fixed non-uniform
mesh was used. Near the newly formed shell this mesh had a
grid spacing less than half the distance moved by the billet
during each timestep. This still allowed for some numerical
mixing downstream of the SWM to simulate relative motion
between the liquid and the shell. At each timestep the
location of the SWM, itself, was determined from the solution
of the POE for the heat transfer model (Eq. (4.3».

6.2 Model for Shell Formation in HCC


A problem with the calculations of both Holleis et al.
(40] and of Pehlke and Ho (41] is that they used experimental
189

data on the final position of the SWM, rather than the heat
transfer model, itself, to determine the location of the
rupture point between the newly grown shell and the moving
shell even in the intermediate stages of the pull phase o£ the
withdrawal cycle. It is also not clear from their work how
the initial position of the SWM and the temperature field
around it were set.
To carry out the simUlation of HCC in an unambiguous
manner, a scenario for shell detachment, growth, breaking and
healing during cyclic withdrawal was added to the present heat
transfer model. It is illustrated in Fig. 42 and described
below.
(i) At the beginning of the pull phase of a new
wi thdrawal cycle, space is created between the
break-ring and the upstream face of the shell
formed in the previous cycle (Fig. 42a). This
space is immediately filled with melt that flows
radially outward. To simUlate this in the model,
at the beginning of the pull phase, the shell, in
effect, moves by two grid spacings along the
x-axis by shifting temperature values. The
temperature at the two columns of grid points
immediately downstream of the break ring are
raised to the melt 'temperature and the velocity
in the longitudinal direction is set to zero.
190

(0)

t (b)

Fig. 42 Scenario for shell detachment, growth, breaking and


healing during cyclic withdrawal for HCC used in the
model: (a) old shell shifted in the x-direction by
transposing temperature values: temperatures at
gridpoints marked by • set to 8~l; (b) a minimum in
shell thickness develops at gr1~ column marked by
arrow; (c) velocities in solid set for next
timestep. For liquid velocities see Fig. 51.
191
(ii) The liquid that fills the space bounded by the
break-ring, the copper mold and the upstream face
of the shell created in the previous cycle
freezes (Fig. 42b). The new shell thus created
has a minimum thickness and hot strength at a
point in between the break ring and the moving
shell. It fractures at this point because of the
wi thdrawal stress. In the simulation at the
timestep following the beginning of the pull
phase, the minimum in shell thickness appears at
the grid column next to the break-ring. The
position of the new SWM, and temperatures around
it are thus defined.
(iii) During the rest of the pull phase the shell
upstream of the SWM remains attached to the
break-ring and that downstream of the SWM moves
away. The space thus created between the fixed
shell and the moving shell is filled by the melt.
This liquid remelts some of the adjacent shell
and then freezes against the relatively cool
copper mold. At each timestep of the simulation
of the pull phase, the position of the SWM is
determined by detecting the location of the
minimum in shell thickness closest to the break
ring. Then the velocity of the shell upstream of
192
the SWM is set to zero while the velocity of the
downstream shell is set to the instantaneous
withdrawal velocity (Fig. 42c). In the absence
of an accompanying momentum equation to simulate
the liquid velocities in this region, the veloci-
ties, temperatures, and properties at each grid
point were set as functions of position and
fraction solid by the method described in
Appendix B.
(iv) During the pause phase of the withdrawal cycle
velocities ever~~here in the casting are zero and
the shell heals (i.e., solidifies) at the SWM and
grows thicker everywhere. Therefore, simulation
velocities are set to zero and, unlike steps (i)
and (iii) above, no modification of temperatures
obtained from the solution of Eq. (4.3) is
needed.

6.3 Input Data and Process Parameters


The effect of varying the withdrawal cycle as well as
melt superheat on the HCC of 114 mm dia billet of 0.7 %C steel
at an average withdrawal speed of 3.5 cm s·1 was investigated.
caster design and operating conditions, described by Miyashita
et al. [18] and discussed in §2.1.3, are given in Table 10.
Thermophysical and solidification properties for the 0.7 %C
193
steel are identical to that given in Table 8. properties for
the BN break-ring were obtained from Touloukian et al. [76].
The scaled-up heat transfer coefficients of Fig. 10 were used
for both the shell/mold and the shell/break-ring interfaces.
The highest value of hgep (0.17 cal.cm· 2 .s·'. ·c·') was used
whenever liquid contacted either the break-ring or the mold.
The fixed refined mesh was described in §6.1. Mesh
openings can l;>e readily deduced from the meshes that are
superimposed on the figures that show the output
(Figs. 43-46). Each cycle time was divided into twenty full
time steps for the calculations.
simulations were carried out for four different
withdrawal cycles, all with the same average casting speed of
3.5 cm· s·' but wi th cycl ic rates ranging from 1 to 2.5
cycles.s·'. These cycles are described in full in Table 11.
The melt inlet temperature was 1530·C for all cases. Cycle 3
was repeated with an inlet temperature of 1515°C to determine
the effect of superheat. It was also run for an average
casting speed of 2.0 cm s·' with melt inlet temperature of
1530°C.

6.4 Results
6.4.1 Effect of cycle Rate
Shell profiles (for fraction solid = 0.05, 0.5, 0.75 and
0.95) downstream of the break-ring are shown in Fig. 43-46 for
194
Table 10
Input Data to Simulate HCC of Fe-0.7%C Steel

Billet Radius: 5.7 em


Casting Speed: 2 • 0 and 3. 5 cm. s-1
Alloy: Fe-0.7%C Steel
Melt Temperature: 1515 and 1530 DC
Thermophysical and Solidification Properties:
Fe-0.7%C Steel: see Table 8
BN break-ring [76]:
p: 27 < 8 < 2000DC, p = 2.2 g.cm- 3

27 < 8 < 327°C, c p = 0.174 + 0.5 X 10-3 8~ cal.g- 1 .oc- 1


327 < 8 < 927 DC, c p = O. 28 + O. 183 x 10- 8
927 < 8 < 1327°C, c p = 0.4114 + 0.417 x 10-4 8
1327 < 8 < 2000°C, c p = 0.475
k: 27 < 8 < 727 DC, k = 0.065 cal.cm- 1 .s- 1 • DC- 1
727 < 8 < 1827 DC, k = 0.0766 - 0.159 x 10- 4 8
Length of Solution Field: 70 cm
Mold Length: 50 cm
Mold Thickness: 1.2 cm
Mold Material: copper (for properties see Table 8)
Heat Transfer Coefficients:
Billet-Mold Gap: see Fig. 10
Mold-Cooling Water: 0.8 cal· cm- 2 • s-1. DC- 1
Mold Air: 0.003 cal cm- 2 • s-1. DC- 1
Billet-Water Spray: 0.02 cal· cm- 2 • 5- 1. Dc- 1
195
Cycles 1'-4, respectively. The withdrawal takes place from
left to right of each figure. In each case profiles in the
vicinity of the break-ring, rather than the whole billet, are
shown at
( i) beginning of pull phase,
(ii) pull at maximum speed,
(iii) end of pull, and
(iv) end of the pause period (i.e., the end of the
cycle).
For Cycle 1 it is seen from Fig. 43a that at the beginning of
the withdrawal stroke a hot zone is created next to the
break-ring by the flow of fresh melt which solidifies into two
parts: the static shell attached to the break ring and the
downstream part attached to the moving shell. As the with-
drawal proceeds the shell thickness in between these two parts
diminishes and during forward pull at maximum speed (Fig. 43b)
there is no solid skin left. However, as the billet starts
decelerating the shell starts growing back (Fig. 43c) and by
the end of the pause period (Fig. 43d) the skin (defined by
fraction solid = 0.95) in this intermediate section grows back
to about 2.0 mm.
The new segment of shell formed next to the break-ring
has pronounced variation of shell thickness o~ waviness along
it. segments formed in the earlier withdrawal cycles do not
196

.. . . . . . . . . .
(0) ::: .. . . . . . . .
::; : : : : : : : : : ; : : : : : : : ; :

11~~;l~~!~~§i
(b)

(c)

.. : : :

(d)

": ": . -.+. . . +. . . +.......+. . . .+.........~ ..........~............

Fig. 43 Computed shell profile for HCC of 114 mm dia.


Fe-0.7% C steel billet using withdrawal Cycle 1 in
'l'able 8. Profiles shown at (a) beginning of pull,
(b) pull at maximum speed, (c) end of pull, (d) end
of pause. At each time the four isotherms shown are
for, from top to bottom, fraction solid = 0.05, 0.5,
0.75 and 0.95. Arrows divide segments from
successive cycles.
197

.. .. ..
(a)r.~~.~~.~~.~.~.~.~~.~··~~~.~~~~.~.~ ... ..

t .
(c) i: . . . ... : : . : . : : : : : . ,

Fig. 44 computed shell profiles for HCC of 114 nun dia.


Fe-O.7% C steel billet using withdrawal cycle 2.
Profiles shown at (a) beginning of pull, (b) end of
pull, (c) end of pause.
198

.. . . .. ...... ..... . . .. . '

Fig. 45 Computed shell profiles for HCC of 114 nun dia.


billet of Fe-O.7% C steel billet using withdrawal
Cycle 3. Shell profiles (a), (b), (c) and (d)
correspond to same stages of the withdrawal cycle as
in Fig. 43.
199

Fig. 46 Computed shell profiles for HCC of 114 mm dia.


billet of Fe-O.7% C steel billet using withdrawal
cycle 4. Shell profiles (a), (b), (c) and (d)
correspond to same stages of the withdrawal cycle as
in Fig. 43.
200
retain such a pronounced waviness which is unlike the
computed results of Holleis et ale [40] or Pehlke and Ho [41].
Instead the qualitative nature of the ~hell profile computed
in this work is closer in appearance to the experimental cast
shell reported by Taylor and Zalner [26] and shown in Fig. 5.
Similar profiles can be seen in Figs. 44-46 for Cycles
2-4. The major differences in the shell profiles obtained for
the four cycles is, of course, because of the differences in
stroke length for each cycle (Table 11).
Despite more than three-fold difference in pause time
between Cycles 1 and 4 (Table 11), it is seen from
Figs. 43d-46d that there was no corresponding difference
between the minimum shell thicknesses at the end of the pause
period. This surprising result may be explained by examining
the computed mold temperatures in the vicinity of the SWM.
The copper mold develops a hot spot when it is exposed to the
melt as the shell ruptures during the pull phase of with-
drawal. A hot spot in the mold can be seen near the entry in
Fig. 47. Although the maximum temperature at the hot spot was
approximately equal for all four cycles, its width increased
as the cycle rate was reduced (Table 12), and the hot spot
delayed the healing of the ruptured shell. Therefore, the
longer pause time in Cycle 1 (0.4 s) did not produce a shell
that was appreciably thicker than that of Cycle 4 with a pause
time of 0.125 s.
Table 11

withdrawal Cycles for HCC simulation

Parameters Cycle 1 Cycle 2 Cycle 3 Cycle 4


Melt Temperature, °c 1530 1530 1530 1530

Avg. Casting Speed, cm· s·1 3.5 3.5 3.5 3.5

Cyclic Rate, cycle· s·1 1.0 1.5 2.0 2.5

Cycle Time, s 1.0 0.67 0.5 0.4

Stroke Length, em 3.5 2.33 1.75 1.4

Peak Speed, em· s·1 8.0 8.0 7.2 6.7

Acceleration, em· s·2 53.7 80.0 96.0 111.6

Acceleration Time, s 0.15 0.1 0.075 0.06

Steady Run Time, s 0.3 0.2 0.175 0.155

Deceleration, em· s·2 -53.7 -80.0 -96.0 -111.6

Deceleration Time, s 0.15 0.1 0.075 0.06

Total Pull Time, s 0.6 0.4 0.325 0.275

Pause Time, s 0.4 0.27 0.175 0.125

f\J
o
....
202

SECONDARY
8t _COOLING

MOLD ~.X
r

Fig. 47 The temperature in the billet and the mold at the


end of the pull cycle during HCC of 114 mm dia.
billets of Fe-0.7% C steel using Cycle 1. Note the
hot spot at the inlet end of the mold/metal
interface.
203
Table 12
Hot spots in the Mold at the End of the Pull Cycle

stroke Peak Mold Length of Hot Zone


Cycles Length. Temperature (Temp. > 350°C)
Cycle s·' cm ·c cm
1 1 3.5 377.9 7.2
2 1.5 2.33 373.7 4.2
3 2.0 1. 75 356.6 3.1
4 2.5 1.4 374.6 3.1

As discussed earlier in §1.1.2.2, the cycle rate affects


the depth of the primary witness mark (PWM). It was found
experimentally that for withdrawal cycles operating at less
than 100 cycles min·', the depth of the PWM rises rapidly. The
present model cannot predict this important result, but it
does produce the expected dependence of shell temperature
adj acent to the break ring on cycle rate (Table 13). As the

Table 13
Subsurface Temperature of Shell Adjacent to the Break Ring
at the End of the Pause Period

Shell Temperature ( °C) at


Distance from Billet Surface
(mm)
Melt
Temp Cycles
Cycle* °C S·1 0 1.75 3.53
1 1530 1 821.2 1123.1 1278.1
2 1530 1.5 832.5 1153.1 1300.2
3 1530 2 829.3 1163.6 1313.1
4 1530 2.5 837.8 1185.9 1328.2
*Average casting speed = 3.5 cm· s·' •
204
cycle rate is increased, the break-ring is exposed to fresh
mel t more frequently so that the temperature in the shell
solidifying next to it also rises. At a depth of 3.53 mm from
the billet surface the shell temperature at the end of the
pause period was 1278°C for Cycle 1 but as high as 1328°C for
Cycle 4. Such a difference in the shell temperature between
the slow and the fast cycles is relatively large when compared
to the 100°C freezing range for the steel. This could
facilitate the remelting and welding of the old shell to fresh
melt in case of Cycle 4, and thereby reduce the depth of PWM.
Results from Table 13 are plotted in Fig. 48 where it can be
seen that the subsurface shell temperature first rises rapidly
with increasing cycle rate and then less rapidly beyond 90
cycles min- 1 • The extent of remelting and welding of the old
shell too will mirror such a variation with cycle rate. If
now the depth of the PWM is assumed to be inversely propor-
tional· to the extent of remelting and welding of the old
shell, then the dependence of the sub-surface shell tempera-
ture on cycle rate seen in Fig. 48 should lead to the
empirically determined dependence of PWM on the cycle rate
(§1.1.2.2) •

6.4.2 Effect of Superheat


Cycle 3 was repeated with ari inlet temperature of 1515°C,
and the shell profiles at various stages of withdrawal cycle
205

0 1200 1330 U
0 o.
Cl
Cl u
u as
1320 't:
~ :J
(I)
:J
(I) 1180
~
~ 1310 0
'G)
0
'G) ,g
,g
E
E 1160 1300 E
E c-)
II)
~
,.. 1290 c-;
!
! 1140 :J
:J
'...
lU 1280 ...Cl
'tii
CD a.
a. E
E ~
{!!. 1120 1270
40 60 80 100 120 140 160
Withdrawal Cycle rate , Cycles min ·1

Fig. 48 Dependence of subsurface shell temperature on


withdrawal cycle rate during HCC of 114 nun dia.
billet of Fe-0.7% C steel.
206
are shown in Fig. 49. No effect of the lS0C reduction in
superheat was found in the shell thickness or profiles at the
SWM during the pull phase (compare Figs. 4Sb,c to 49b,c). An

examination of temperatures above the SWM showed less than 3 °C


difference between the two cases. with an inlet temperature
of lS30°C (Fig. 4Sb,c), the profile for fraction solid of O.S
and 0.9S showed some waviness in the second segment of the
shell (formed during the last but one withdrawal stroke).
This waviness almost disappeared from the profiles for the
inlet temperature of lSlS0C (Fig. 49b,c), which were moreover
thicker by about 1 mm compared to the corresponding profiles
for the higher inlet temperature. From these results it
appears that the effect of lowering the lower superheat does
not propagate all the way down to the mold wall but is strong
closer to the billet centerline. At the exit end no effect of
the lower superheat can be seen in the profiles. This can be
ascribed to the rate limiting effects of the low mold-metal
heat transfer coefficient and longer residence time of the
shell far from the break-ring (see Fig. 10).

6.4.3 Effect of Casting Speed


Cycle 3 was repeated with an average casting speed of 2.0
cm.s·', a peak forward speed of 4.1 cm·s·' and a stroke length
of 1.0 cm, keeping all other parameters same as in Table 11;
the results are shown in Fig. so. Again comparing with
207

(0) .: :. :. :. .: .: .: ;. .: :. : : : ... : : : :
.
:
.
: : : . i

]1;·!IflIm::1~El~tl=:t~~J:~~~ft:t-l ... '.' ...••: .....•.-....•;.:" ..... ~.:" ...... ~...... ~.:•..•..• :.'• ••...•.. ~..•...•••:..•..•.•...: ......... ~.; ...... ~.•.....•....1
; . : . : :: : : 1

, .. .,

Fig. 49 Computed shell profiles for HCC of 114 mm dia.


billet of Fe-0.7% C steel using Cycle 3 and a melt
inlet temperature of 1515°C. Shell profiles (a),
(b), (c) and (d) correspond to the same stages of
the withdrawal cycle as in Fig. 43.
208

(0)

(b) ..i. ~ ...~ ..i...!...~ ...~....~ ...i....~.....:.....:....~.............;......L.....;.......t.......~ ........:........).........;.........;..........:..........~ ..........

i;~~!ir~:!·i:~"=)r;(~~:~~:?~~~~!·~;
(c)
.:'.:·;.···c.:•• ·.: :;·.: L·~: : : :i'.·.·.·.·~.· ·: .: •• ~::::~·..·.·.·.t: :·:~.: : :!: : : L:·.·.:.·.: : : : :~: : : :~:.: .. ::::;.::. :.,,::.;::.:::::::;:::::::... :::: . :: .
.......... . ...... ; ...... :........ :. ...... ,'...... ": ... ,, ....:....';' .,
: ;.... ~... ,":. " ... .. "' .... ... ..... ; .... :.. .. .

(d) . ;.. ~...:... ~...:.,,~...~....:...;....... ,,:..... ~ .....:.............;...... :. ......;.......L...... :. ........ ...................i......... ; ". '" ........................

JI;J~.;;;~:.~";~:;;:~:";~::;.:
Fig. 50 Computed shell profiles for HCC of 114 mm dia.
billet of Fe-O. 7% C steel using cycle 3 but an
average velocity of 2.0 cm s·'. Shell profiles (a),
(b), (c) and (d) correspond to the same stages of
the withdrawal cycle as in Fig. 43.
209

Fig. 45, it is seen that at intermediate distances from the


break-ring (2 to 3 segment lengths) the solidus is thicker by
about 1.6 mm. It is to be noted that this intermediate zone
is downstream of the hot spot in the mold for this case.
Because the slower withdrawal rate results in a longer
residence of the shell near the break-ring, where the mold-
metal heat transfer coefficient is highest and the mold is at
a relatively low temperature, the shell becomes thicker during
the pull phase.

6.5 Conclusion
The scenario of shell formation described in §6.2 was
incorporated into the heat transfer code CASTSIM-PC. The
resulting simulation reproduced the characteristic wavy shell
profile in HCC. The simulation was done without recourse to
partitioning of the newly formed shell between the static and
moving parts on the basis of empirical data on the position of
the secondary witness mark, as had been done in previous
attempts [40,41].
A critical part of this modelling effort was the simul-
taneous solution of temperature fields in the billet, mold and
break-ring separated by interfaces. This was made possible by
the development of the numerical techniques and equations
discussed in chapters 3 and 4. Unlike previous modelling of
HCC [41], it was no longer necessary to assume a uniform mold
210

temperature or a constant mold-metal heat transfer coefficient


along the length of the mold. By solving for temperatures in
both the mold and the casting, it was possible to detect a
transient hot spot in the mold. For the slower withdrawal
cycles, this hot spot deterred the healing of the ruptured
shell during the pause period. Mold designs that can prevent
this hot spot or cool it rapidly during the pause cycle will
be beneficial to healing at the SWM and allow faster casting
rates.

6.6 Future Work


Equipped with this model and program, practitioners of
horizontal continuous casting operations will be able to find
many other ways to study and improve the process. As an
example the effect of thermal barrier coatings at the triple
point in the break-ring/mold assembly on the depth of primary
witness marks could be studied indirectly through its effect
on sub-surface shell temperatures. other improvements may
also be initiated from simUlation results. However, in order
to simUlate the effect of cycle rate on the formation and
depth of the primary witness marks, it will be necessary to
model the fluid flow, solidification and melt-shell wetting in
the gap between the break ring and the moving shell. In order
to carry out the computations efficiently, it would be
desirable to also evaluate and/or develop numerical methods,
211

e.g. multigrid techniques [77,78], for interpolating solutions


between micro and macro level domains, as well as adaptive
grids [79].
212

APPENDIX A

MESH REFINEMENT BY COORDINATE TRANSFORMATION

A coordinate transformation may be achieved by several


ways. The well known method for body fi tted coordinates
involving Jacobians [63] that can transform an arbitrary field
into a rectangular mesh is, however, unsui table for this
problem with internal interfaces. The coordinate transforma-
tion method used involves transforming a non-uniform rectangu-
lar mesh into a rectangular uniform mesh. The refinement was
carried out for both x and r directions in cylindrical
coordinates.
The relation between the original and the transformed
coordinates is expressed as a polynomial, and by specifying
sufficient number of constraints the coefficients of the
polynomial can be solved. The procedure is illustrated with
a simple one dimensional problem.
Consider a field 0 ~ x ~ b, in which the mesh is fine in
the interior at x = xI and coar:s;e near the boundaries. After
transformation to the X coordinate, the new mesh will be
uniform and the point x = xI will be located at X = XI. Six ~e

x = 0 at X = 0, let the original coordinate x be related to


213

the transformed X coordinate by the following third order


polynomial:

(A.1)

where co' c, and c 2 are constants.


If Xi = i~x where i = 0,1, • ,n, then ~X = X,/n
and

(A.2)

(A.3)

(A.4)

For the non-uniform mesh, ~xi = xi - x i _,.


At x = xi the desired condition is d(~x)/dX = O. Now we
show that if d 2x/dX2 = 0, then d(~x)/dX = 0 as well. By
definition

~x c, (X + ~X)2 + '3
- Co ex + ~ X) + 2" C2 ex + ~X)3

(A.5)
214
Then Eg. (A.5) is expanded and resulting terms with (AX)2 and
(AX)3 are neglected for small AX. The result is

(A.6)

Recall that AX is constant. Then

(A.7)

By comparing Egs. (A.4) and (A. 7), the conclusion is that


d(AX)/dX = 0 if d 2x/dX2 = o.
Now consider a mesh with n nodes. At m, nodes, values of
the original coordinate x and transformed coordinate X are
specified. At m2 other nodes values of x, d 2x/dX2 and X are
given. These m2 nodes are special in that they are located
where it is desirable to have either minimum values of AX or
maximum values of AX. Then a pOlynomial of order n = m, + 2m2
describes the transformation. If in addition at x = X = 0
values of dx/dX (= a.,) and d 2x/dX2 (= a o) are also specified
then the transformation from x to X is described by the
following set of polynomials of order n+2.

x _ a ..1 X + a o X2 + a, x3 + • • • + • aJ Xj+2 + • • • +
2 3 J + 2
215

dx
dX - a -, + a 0 X + a , Xl + •• • + aJ xJ +' + ••• +anXn+'

or

a, X3 + •••
-3 aJ
+..,-~
j + 2
XJ+l + ••• - x - X [a_,. + :0 X]

(A.8)

a1 X l + ••• + aj XJ+' + ••• an Xn+1 - X' - (a_, + aox)


(A.9)

2a 1 X + ••• + (j + l)ajxj + ••• + (n + l)a nXn - x" - ao


(A.10)

where the prime and double prime represent first and second
derivatives, respectively. There are n = m, + 2mz unknown
coefficients a 1 . • • an' For the coordinate transformation
problem equations A.8 and A.10 are used. There will be m, + mz
of Eq. (A.8) with known (x,X) values as well as mz of
Eq. (A.10) with known (x, dlx/dXl) values resulting into n =
m1 + 2ml equations to solve.
216
The problem can be set into the following matrix equation

xa-b (A.ll)

where X is a n x n matrix consisting of the coefficients of


the left hand side of Egs. (A.8)-(A.10), the n x 1 solution
vector a contains the unknown a J and the n x 1 vector b the
known values on the right hand side. The matrix equation can
be solved in order to obtain aJs.
217

APPENDIX B

LIQUID VELOCITIES AND TEMPERATURES


NEAR THE SECONDARY WITNESS MARK IN HCC

with the aid of Fig. 51, the liquid flow that occurs in the
region between the fixed and the moving shell of the HCC
process during the withdrawal cycle may be qualitatively
explained as follows. This flow is driven by the space
created by the movement of the moving shell away from the
fixed shell. At the tips of the two shells near the mold wall
the flow is constricted in the radial direction by the
dendrites in the respective mushy zones. Away from the mold
wall the mushy zone of the fixed shell allows flow and
consequently the flow there will have small longitudinal
components. Further away from the mold wall the flow will be
mostly longitudinal. Calculations were done in which all
radial flow of the melt was ignored; these calculations
resulted in shells that were unrealistically thick with a
depression at the SWM that was hardly perceptible. Assuming
radial flow everywhere in the newly forming shell, ot:! the
other hand, produced very thin shells. Therefore to arrive at
an acceptable shell profile, it was necessary to use a
realistic flow pattern adjacent to the shell.
Because this work does not include momentum equations
required to simulate the flow described above, it was
r r, , ,_
III/rllfll1
x

I
I _~_I_~
I I__
I _ I 't
'---
I BREAK
-1- -RING
/'t;~~
1 I L
-I- -r-I
I I ..L
-I hb"t- I
I I -1. -I-¥J-
-I - , - I
-~ _l_t-
I 1 1
-,-f--r
II '/ III I II

MOLD

Fig. 51 Proposed velocity pattern in the liquid adjacent to the break-ring


and the shell during the pull phase of HCC. N
....00
219
necessary to approximate the velocity field in and near the
shell to make a realistic accounting of advective terms in the
energy equation. This was done in the following manner.
Suppose over a time interval ~t during simUlation the
longitudinal velocity is Vx then the distance between the
fixed and the moving shell at the end of the timestep will be

(B.1)

The volume of liquid in an element of sides t and h f and unit


depth located between the fixed and the moving shell is

v, = t hb (B.2)

where hb is the height of the break-ring. The volume of


liquid in an element of sides ~x, and ~rj and unit depth
located in the mushy zone is

v, = 8X, 8rj (1 - fs) (B.3)

where fs is the fraction solid at that element. Now if the


volume of liquid flowing out to adjacent elements, in order to
fill the space created by withdrawal is V2 then the total
volume of liquid flowing into the element is

v = V, + V2 (B.4)
220

In that case the maximum radial velocity of the liquid melt


(with no longitudinal flow) at a given gridpoint (i,j) is

(B.5)

At most gridpoints the liquid will have both radial as well as


longitudinal velocities. At such a gridpoint in that case the
radial velocity of the liquid is defined by

v ra = cfd vm (B.6)

and the longitudinal velocity of the liquid is

(B.7)

where the coefficient c fd determines the extent of the radial


velocity depending on d J , the radial distance from the mold
wall and fs' the fraction solid. Somewhat arbitrarily we
define the flow direction coefficient

(B.8)

If the calculation of velocities is started from the tip of


the moving shell adjacent to the mold wall (where there is no
liquid outflow) and continued to locations upstream, then both
v xa and v ra in this region will be uniquely determined. The
volume V2 will be calculated during this procedure by adding
the appropriate radial and longitudinal inflows into adjacent
locations (i.e. gridpoints).
221

After the local velocities are determined, the


temperatures can be updated by assuming complete equilibration
between the existing liquid/solid at an element and fresh melt
from adj acent upstream locations. The proportion between the
two is (1 - V2/V1 ) and V2/V1 • For locations within the mushy
zone of the fixed shell, the effect of the static solid is
taken into account to obtain the new temperature with the
assumption of full mixing. In the region immediately sur-
rounding the tips of the fixed and moving shells at the mold
wall, the radial velocities can be so high that the distance
travelled by the liquid over a typical timestep exceeds the
fine grid spacing in the radial direction. In such cases the
updating of temperature described above needs to take into
account temperature at several adjacent gridpoints and the
simple procedure for updating temperatures involving only
adjacent gridpoints becomes invalid. Therefore, the effect of
the radial component of velocity on the local temperature is
implicitly accounted for by the use of an enhanced thermal
conductivity k. Following Michalek and Dantzig [35] for the
largest radial velocities (near the tip of the two shells), k
in the radial direction (in coefficients a 3 and a 4 of Eqs.
(4.17) or (4.19), respectively, is multiplied by 3.5 and by
1.0 for the smallest radial velocity. Multiplication factors
for intermediate radial veloci ties is determined by
interpolation.
222

REFERENCES

1. Y. S. Touloukian, Thermophysical Properties of High Temperature Solid


Afmeri~,vol. 1, pp. 7-18, MacMillan, New York, 1967.
2• Y. S. Toulouldan, Thermophysical Properties of High Temperature Solid
Afmeria~, vol. 1, pp. 622-633, MacMillan, New York, 1967.

3. W. J. Bergmann, Solidification in Continuous casting of


Aluminum, Afetall Trans., vol. 1, pp. 3361-3364, 1970.
4. S. E. Taylor, TechnicaGuss, pvt. comm., 1986.
5• Y. S. Toul oukian, Thermophysical Properties of High Temperature Solid
Afateria~, vol. 1, pp. 578-591, MacMillan, New York, 1967.

6. ' Y. S. Touloukian, Thermophysical Properties of High Temperature Solid


Afateria~, vol. 1, pp. 694-705, MacMillan, New York, 1967 •

. 7. I. V. Samarasekera and J. K. Brimacombe, The Influence of


l-Iold Behavior on the Production of Continuously Cast
Steel Slabs, Af~an Trans.B, vol. 13B, pp. 105-116, 1982.
8. D. J. Harvey, G. L. Vaneman and F. J. Webbere, Horizontal
continuous Casting of Nickel Base Alloys, in F. Mollard
and Y. Murty ( ed • ), Continuous Casting of Small Cross Sections,
pp. 157-171, TMS-AIME, Warrendale, PA, 1981.
9. H. A. Krall and H. Huber, Design Characteristics of
Horizontal continuous Casting Plants for Production of
Square Billets, Metallurgical Plant and Technology, vol. 6,
pp. 44-60, 1983.
10. H. A. Krall and B. Schmitz, Some Factors Affecting High
Yield and Quality when Casting Speciality Steel Billets
on Horizontal continuous Casting Plants, Metallurgical Plant
and Technology, vol. 8, pp. 40-47, 1985.
11. P. Voss-Spilker and W. Reichelt, A Review of Horizontal
continuous Casting of Metals with Special Reference to
Steel, Metals Forum, vol. 7" pp. 79-97, 1984.
223

12. P. Voss-Spilker and W. Reichelt, A Horizontal continuous


caster for High Alloy Steel Grades and Sections Upto 14
inches Round, Iron and Steelmaker, vol. 13, pp. 28-33, 1986.

13. R. Hentrich, D. L. Sharma, D. Dittert and E. Roller, The


Horizontal Casting of High-Alloy Steels using the
Krupp oscillating Mold Process, Proc. 43rd Electric Arc Furnace
Conference, TMS-AIME, Warrendale, PA, pp. 34-41, 1985.
14. R. Hentrich, D. Detleff, R. Roller and M. Buch, The
KRUPP Horizontal Caster a Production Tool for the
Electric Furnace Mel tshop, Proc. 44th Electric Arc Furnace
Conference, TMS-AIME, Warrendale, PA, pp. 133-141, 1986.
15. P. Machner, E. Reisl, R. Tarmann, G. Holleis and W.
Polanscheutz, Horizontal Continuous Casting Technology
for Special Steels, Proc. 43rd Electllc Arc Furnace Conference,
TMS-AIME, Warrendale, PA, pp. 15-18, 1985.

16. M. Haissig, Horizontal Continuous Casting - A Technology


for the Future, Iron and Steel Engineer, vol. 61, pp. 65-70,
1984.
17. M. Haissig, Horizontal Continuous Casting in Mini Mills,
IronandSteelmaker, vol. 12, pp. 359-367, 1985.
18. Y. Miyashita, S. Miyahara, Y. Ueno, M. Ishikawa and A.
Honda, Development and Metallurgical Analysis of the
Horizontal continuous Casting, Iron and Steelmaker, vol. 8,
pp. 22-28, 1981.

19. T. Koyano and M. Ito, Development and Industrialization


of HORICAST - NKK's New Horizontal continuous Casting
Process, Ironmaking and Steelmaking, vol. 10, pp. 289 - 295,
1983.
20 • S. Tanaka, K. Taguchi, A. Honda, T. Kawawa, E. Sunami and
S. Komori, Development of Horizontal continuous Caster
for steel Billets, Tr~ISIJ, vol. 24, pp. 973-982, 1984.

21. D. Toothill, Application of HORICAST for continuous


Casting of stainless and High Alloy Steels, Ironmakingand
Steelmaking, vol. 15, pp. 43-48, 1988.
224
22. K. Nakai, Y. Umeda, Y. Sugitani and M. Miura, Effects of
Factors on the Stability of casting in Horizontal
Continuous Casting, TransISII, vol. 24, pp. 983-991, 1984.

23. T. Sakane, H. Kodama, H. Kiyoto, Y. Sugitani, S. Hiraki


and K. Nakai, Production of stainless Steel and High
Nickel Alloys by the Horizontal continuous casting
Process, TransISII, vol. 24, pp. 19-23, 1984.

24. A. J. Zalner, Horizontal continuous Casting - The Second


Generation, Fachberichte Huttenpraxis Metallweiterverarbeitung,
vol. 23, pp. 604-606, 1985.
25. A. J. Zalner and S. E. Taylor, Horizontal continuous
casting of stainless Steel at Armco's Baltimore Works,
Iron and Steel Engineer, vol. 62, pp. 37-44, 1985.

26. S. E. Taylor and A. J. Zalner, TechnicaGuss, Horizontal


Casting of Specialty Steel Alloys, presented at
Convention of Iron and Steel Engineers, 1985.
27. T. Wada, M. Suzuki and T. Mori, High Speed casting of
3 meters/minute on the NKK Fukuyama Works' no. 5 Slab
Caster, Iron and Steelmaker, vol. 14, pp. 31-38, 1987.

28. D. Gupta and H. P. Cheskis, Composite Mold for continuous


Thin strip casting, u.S. Patent no. 4,773,469, Sept. 27,
1988.

29. M. C. Flemings, Solidification Processing, McGraw-Hill, New York,


pp. 11-12, 1974.
30. D. J.-P. Adenis, K. H. Coats and D. V. Ragone, An
Analysis of the Direct-chili-Casting Process by Numerical
Methods, 1 oflnst. of Metals , vol. 91, pp. 395-403, 1962-63.

31. E. Miziker, Mathematical Heat Transfer Model for


Solidification of Continuously Cast steel Slab, Trans.Met
SocAIME, vol. 239, pp. 1747-1753, 1967.

32 • R. J. Dippenaar, I. V. Samarasekera and J. K. Brimacombe,


Mould Taper in Continuous-Casting Billet Machines, Proc
43rd Electric Arc Furnace Conference, TMS-AIME, Warrendale, PA,
pp. 103-117, 1985.
225

33. D. C. Weckman, R. J. Pick and P. Niessen, Numerical Heat


Transfer Model of the D.C. continuous casting Process, Z.
Metal/kunile, vol. 70, pp. 750-757, 1979.
34. D. C. Weckman and P. Niessen, A Numerical simulation of
the D.C. continuous Casting Process Including Nucleate
Boiling Heat Transfer, Metal! Trans. B, vol. 13B,
pp. 593-602, 1982.
35. K. E. Michalek and J. A. Dantzig, Modelling of In-Mold
Heat Transfer in Continuous Casting of Steel, Proc. Conf. on
Modelling and Control of Casting and Welding Processes, TMS-AIME,
Warrendale, PA, pp. 497-516, 1986.

36. M. Bamberger and B. Prinz, Mathematical Modelling of the


Temperature Field in continuous Casting, Z. Metal/kunile,
vol. 77, pp. 234-238, 1986.
37. E. Selcuk, Mathematical Modelling of Heat Transfer in
Horizontal continuous Systems, Proc. Conf. on Solidification
Procesdng, Institute of Metals, London, pp. 339-342, 1987.
38. J. P. Verwijs and D. C. Weckman, Influence of Mold Length
and Mold Heat Transfer on Horizontal continuous Casting
of Nonferrous Alloy Rods, Metal! Trans B, vol. 19B,
pp. 201-212, 1988.
39. D. R. Croft, D. Toothill and B. Telford, Heat Transfer in
Horizontal continuous Casting, in Numerical MethDds in Thennal
Probkms, Proc. Third International Conference, vol. 2,
pp. 1365-1376, Pineridge Press, Swansea, Wales, 1985.
40. G. Holleis, K. H. Schwaha, P. Machner and R. Tarmann,
Investigations on the Fundamentals of Shell Formation in
Horizontal continuous Casting, Proc. 43rd Electric Arc Furnace
Conference, TMS-AIME, Warrendale, PA, pp. 349-357, 1985.
41. R. D. Pehlke and K. Ho, simulation of Horizontal
continuous casting with Incremental Strand Movement, Proc.
43rd Electric Arc Furnace Conference, TMS-AIME, Warrendale, PA,
pp. 119-129, 1985.
42. T. M. Shih and Y. N. Chen, Comparison of Finite
Difference Method and Finite Element Method, in T. M.
Shih ( ed. ), Numerical Properties and Methodologies in Heat Tran.sfer,
Hemisphere Publishing Corp., New York, pp. 33-54, 1981.
226
43. S. N. singh and K. E. Blazek, Heat Transfer and Skin
Formation in a continuous casting Mold as a Function of
Steel Carbon Content, 1. of Metals, vol. 26, pp. 1-9, 1974.

44. K. E. Blazek, Mold Heat Transfer During continuous


Casting - Part I, IronandSteelMaker, vol. 14, pp. 49-50,
1987.
45. K. E. Blazek, Mold Heat Transfer During continuous
Casting - Part II, Iron and Steel Maker, vol. 14, pp. 36-38,
1987.
46. K. E. Blazek, Mold Heat Transfer During Continuous
Casting - Part III, IronandSteelMaker, vol. 14, pp. 42-44,
1987.
47. K-O Yu, Comparison of ESR-VSR Processes - Part I, Heat
Transfer Characteristics of Crucible, Proc. Vacuum Metallurgy
Conference, Iron and Steel Society of AIME, Warrendale, PA,
pp. 83-92, 1985.
48. P. M. Auman, D. K. Griffiths and D. R. Hill, Hot strip
Mill Runout Table Temperature Control, Iron and Steel Engineer,
vol. 44, pp. 174-179, 1967.
49. M. Larrecq, J. P. Birat, C. Saguez and J. Henry,
optimization of casting and Cooling Conditions on Steel
Continuous Casters - Implementation of optimal Strategies
on Slab and Bloom Casters, in Applic. of Math. and Phys. Models in
the Iron and Steel Industry, I&SS 3rd Process Tech. Conference,
Pittsburgh, PA, vol. 3, pp. 273-282, 1982.
50. M. Bamberger and B. Prinz, Determination of Heat Transfer
Coefficients During Water Cooling of Metals, Materials Science
and Technology, vol. 2, pp. 410-415, 1986.

51. N. Lambert and M. Economopoulos, Measurements of Heat


Transfer Coefficients in Metallurgical Processes, 1. of the
Iron and Steel Institute, vol. 208, pp. 917-928, 1970.

52. B. G. Thomas, Application of Mathematical Models to the


Continuous Slab casting Mold, Trans. IS&S, Iron and Steelmaker,
vol. 16, pp. 53-66, 1989.
53. H. Sundquist and G. Veronis, Simp~e Finite-Difference
Grid with Non-constant Intervals, Tellus, vol. XXII,
pp. 27-:H, 1970.
227
54. W. C. Thacker, A Brief Review of Techniques for
Generating Irregular computational Grids, Int. 1. for Numerical
Methods in Engineering, vol. 15, pp. 1335-1341, 1980.
55. J. A. Sills, Transformations for Infinite Regions and
their Application to Flow Problems, AUAAJounud, vol. 7,
pp. 117-123, 1969.
56. T. Gal-Chen and R. C. J. somerville, On the Use of a
Coordinate Transformation for the Solution of the
Navier-Stokes Equations, 1. of Compo Phys., vol. 17,
pp. 209-228, 1975.
57. B. Carnahan, H. A. Luther and J. o. Wilkes, Applied
Numerical Methods, pp. 462-463, John Wiley, New York, 1969.

58. W. F. Ames, Numerical Methods for Partial Differential Equations, 2nd


ed., pp. 230-238, Academic Press, New York, 1977.
59. H. Motz, The Treatment of Singularities of Partial
Differential Equations by Relaxation Methods, ~art.AppL
Maths., vol. 4, pp. 371-377, 1946.
60. G. E. Bell and J. Crank, A Method for Treating Boundary
Singularities in Time-Dependent Problems, 1. Inst. Maths.
Applies., vol. 12, pp. 37-48, 1973.

61. G. E. Bell and J. crank, Simple Finite Difference


Modification for Improving Accuracy Near a Corner in Heat
Flow Problems I Int. 1. for Numerical Methods in Engineering, vol. 10,
pp. 827-832, 1976.
62. F. C. Trutt, E. A. Erdelyi and R. F. Jackson, The Non-
Linear Potential Equation and its Numerical Solution for
Highly satu~ated Electrical Machines, IEEE Trans. Aerospace,
vol. 1, pp. 430-440, 1963.
63. J. Crank, Free and Moving Boundary Problems, pp. 163-282,
Clarendon Press, Oxford, 1984.
64. N. Shamsundar and E. M. Sparrow, Analysis of
Mul tidimensional Conduction Phase Change via the Enthalpy
Model, Trans. ASME-1. Heat Transfer, vol. 97, pp. 333 - 340,
1975.
228
65. R. M. Furzeland, A comparative study of Numerical Methods
for Moving Boundary Problems, 1. Inst. Maths. Applies., vol. 26,
pp. 411-429, 1980.

66. G. D. Smith, Numerical Solution of Partial Differential Equations, 2nd


ed., pp. 37-40, Clarendon press, Oxford, 1978.
67. A. M. Winslow, Numerical Solution of the Quasilinear
Poisson Equation in a Nonuniform Triangle Mesh, 1. ofComp.
Phy~, vol. 2, pp. 149-172, 1967.

68. I. Ohnaka and K. Kobayashi, Flow Analysis During


Solidification by the Direct Finite Difference Method,
Trans. lSI!, vol. 26, pp. 781-789, 1986.

69. H. J. Carslaw and J. C. Jaegar, Conduction of Heat in Solids,


2nd ed., pp. 99-102, Oxford University Press, Oxford,
U.K., 1959.

70. M. N. Ozisik, HeatConduction, p. 440, John Wiley, New York,


1980.
71. M. C. Flemings, Solidification Processing, pp. 33-36, McGraw-
Hill, New York, 1974.

72 • L. Lapidus and G. F. Pinder, Numerical Solution of Partial


Differential Equations in Science and Engineering, pp. 170-211 , John
Wiley, New York, 1982.

73 • R. Peyret and T. D. Taylor, Computational Methods for Fluid Flow,


pp. 37-44, springer Verlag, New York, 1983.
74. D. B. Spalding, A Novel Finite Difference Formulation for
Differential Expressions Involving Both First and Second
Derivatives, Int. 1. for Numerical Methods in Engineering, vol. 4,
pp. 551-559, 1972.

75. I. Gladwell and R. Wait, A Survey of Numerical Methods for Partial


Differential Equations, pp. 197-211, Oxford University Press,
New York, 1979.

76. Y. S. Touloukian, Thennophysical Properties of High Temperature Solid


Materials, vol. 5, pp. 499-508, MacMillan, New York, 1967.
77. R. E. Phillips and F. W. Schmidt, Multigrid Techniques
for the Numerical Solution of the Diffusion Equation,
Numer. Heat Transfer, vol. 7, pp. 251-268, 1984.
229
78. R. E. Phillips and F. W. Schmidt, Multigrid Techniques
for the Solution of the Passive Scalar Advection-
Diffusion Equation, Numer. Heat Trans/er, vol. 8, pp. 25-43,
1985.

79. J. F. Thompson, Z. U. A. Warsi and C. W. Mastin, Numerical


Grid Generation, pp. 1-32, North-Holland: Elsevier Science
Publ. Co., New York, 1985.

Вам также может понравиться