Anda di halaman 1dari 20

International Journal of Thermal Sciences 86 (2014) 68e87

Contents lists available at ScienceDirect

International Journal of Thermal Sciences


journal homepage: www.elsevier.com/locate/ijts

Modeling of 3-D turbulent transport phenomena and solidication of


a direct chill caster tted with a metallic-foam-plated combo bag
Latifa Begum, Mainul Hasan*
Department of Mining & Materials Engineering, McGill University, M.H. Wong Building, 3610 University Street, Montreal, QC H3A 0C5, Canada

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 20 August 2013
Received in revised form
16 June 2014
Accepted 24 June 2014
Available online

The 3-D numerically simulated steady state results of an industrial size vertical Direct Chill (DC) slab
caster tted with a metallic-foam-plated combo bag melt distributor for AA-1050 aluminum alloy are
reported. The turbulence in the melt and the mushy region solidication of the alloy are modeled
through the popular low Reynolds number ke model of Launder and Sharma and the enthalpy-porosity
scheme, respectively. The transport of melt through the foam is modeled using the BrinkmaneForchheimer extended Darcy equation. The verication of the numerical model is performed by comparing the
predicted and measured solidication front for AA-3104 rolling ingot reported in the literature. The inlet
melt superheat and the porosity of the metallic foam of the combo-bag were kept xed at 32  C and 0.9,
respectively. Parametric studies are carried out by varying two important parameters of the process, viz.,
the casting speed and the heat transfer coefcient (HTC) at the metal-mold contact region. Specically,
the casting speed and the HTC are varied from 40 to 100 mm/min and from 750 to 3000 W/(m2-K),
respectively. With the increase of the casting speed, the solidication process is delayed while the solidshell thickness increased with the increase of the HTC. The predicted results are presented for the solidshell thickness, the sump depth, the mushy thickness, the surface temperature as well as the nondimensional turbulent viscosity contours. In addition the temperature elds and velocity proles with
streamlines are also provided.
2014 Elsevier Masson SAS. All rights reserved.

Keywords:
Modeling
Mushy-zone solidication
3-D DC casting
Rolling ingot
Aluminum alloy
Turbulent melt
Metallic foam

1. Introduction
The vertical direct chill casting, often referred to as DC casting
(DCC), is a semi-continuous casting process which is a relatively
efcient and economical process for casting large size rolling ingots
and round extrusion billets for nonferrous metals and their alloys.
Among the non-ferrous metals, aluminum is the most extensively
used metal spanning a wide range of applications including, in
aerospace, automotive, electronic, construction, etc. More than 90
pct of the large slabs and billets of aluminum alloys are nowadays
cast using this technology [1]. The cast slabs are rolled into plates,
sheets, strips and foils for various end uses. The round billets are
usually extruded in the form of bars, rods, pipes and tubes for
diversied practical uses. In this process, molten metal is poured
into a water cooled mold whose open bottom is initially kept closed
with a bottom block such that the mold wall(s) and the base bottom

* Corresponding author. Tel.: 1 514 398 2524, 1 514 924 8542 (mobile); fax: 1
514 398 4492.
E-mail addresses: Latifa.begum@mail.mcgill.ca (L. Begum), Mainul.hasan@
mcgill.ca, mnlhsn73@gmail.com (M. Hasan).
http://dx.doi.org/10.1016/j.ijthermalsci.2014.06.031
1290-0729/ 2014 Elsevier Masson SAS. All rights reserved.

block forms a mold cavity. Metal is delivered in a controlled manner


using a suitable melt delivery system into the mold where primary
cooling of the melt takes place. Once a solid shell is formed on the
mold wall(s), the bottom block is slowly lowered by hydraulic
cylinders. The ingot at this point embodies liquid metal inside its
cavity. When the partially solidied ingot is out of the mold, it is
further cooled by water jets coming out from the openings at the
mold bottom and this cooling of the semi-solid ingot is called the
secondary cooling. Depending on the aspect ratio of the ingot and
casting speed, within 0.5e1 m length the process reaches steady
state [2]. Once the cast is of sufcient length (usually 8e10 m [2]),
the casting is stopped and the inlet melt delivery is suspended.
After that the ingot is taken out from the casting pit with a hydraulic system. About 85e90% of heat is removed in the secondary
cooling region while only 10e15% of heat is removed through the
primary cooling in the mold [3].
Although, the above description of the DCC process appears to
be quite simple, unfortunately a number of factors, e.g. alloy
composition, solidication range of the alloy, added grain rener,
aspect ratio of the caster, casting speed, inlet melt superheat,
cooling water ow rate, temperature in the mold, etc. play

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

interacting roles and control the solidication of the metal and


hence the yield and quality of the casts. Among the various defects
in the cast, some are internal others are surface related. The surface
cracks constitute, starting cracks, edge/face cracks, stress/midway
cracks, etc. while internal defects arise due to macro- and microsegregation, micro-voids, hot tears and cold cracks, and distortions in the ingot [4e7]. As a result, each casting facility, through
trial and error and from past experience usually xes the casting
process parameters to obtain an acceptable quality cast.
Many of these defects in the cast are believed to be related to the
melt ow patterns in the sump. In order to control the inlet melt
ow to the mold one approach is to use a suitable melt feeding
system. In this regard the combo bag is in use since 70s and
nowadays it is almost universally used in aluminum slab DC casting
processes [8]. Although, one-time use of the deformable porous
combo bag, which is usually made of berglass, is in common use in
this industry for melt delivery, unfortunately there is hardly any
serious attempts taken to date to model such a system in the
presence of the solidication process. An extensive search of the
coupled turbulent ow and solidication literature did not reveal
any studies which have dealt with the complex problem of turbulent convective heat transfer and solidication in the presence of a
porous lter. Because of the lack of a similar study in the literature
which has considered the combo bag with a porous bottom plate as
well as solidication, hence only the works which are available in
the open literature and have some peripheral relevance to the
present study is briey reviewed below.
Before we start the literature review it should be mentioned rst
that combo bag is a standard melt distributor in majority of the DCC
facilities operating around the world. One of the major suppliers of
combo bag is the PYROTEK, INC., Saguenay, Quebec, Canada. Most of
the theoretical and experimental studies concerning combo bag
system for DCC process were carried out by various DCC industries
and academic institutions in cooperation with the personnel of
PYROTEK, INC. Since combo bag, which is usually made of berglass
fabrics, needs to be changed before each cast, hence to reduce costs
Tremblay and Lapointe [9] proposed a new reusable rigid distributor for DC sheet ingot casting and called it as Reusable Molten
Aluminum Distributor (ReMAD). The latter system consists of a
rigid open structure which tightly holds the porous berglass bag.
Through water modeling studies the authors showed that REMAD
resulted in a ow at the free surface that was much slower
compared to the standard combo bag. It is to be recognized here
that a slow ow eld at the melt free surface has a positive impact
on the quality of the cast because less oxide inclusions and broken
oxide lms will be transported from the top free surface into the
cast.
Since the standard berglass combo bag is exible in nature,
hence it undergoes unpredictable distortions during the actual cast
operations. In order to incorporate the inuence of the distortion
on the melt ow, Kieft et al. [10] adopted a new strategy in their
modeling of the exible combo bag in melt feeding system. The
reported work was divided into two parts. In the rst part, full-scale
water modeling experiments were performed to investigate the
ow behavior for a distorted combo bag. At the exible distribution
bag outlets (side and bottom openings), the ow was measured
quantitatively by using the image analysis technique PIV (Particle
Image Velocimetry). In their modeling part of the work the authors
used this measured velocity eld at the exit of the combo bag. They
modeled the 3-D ow eld using the commercial CFD software
package CFX version 4.4. The authors only predicted the velocity
eld without solving for the thermal eld. Moreover, the authors
did not compare their measured and predicted velocity elds with
any published results. In reality, the deformation of the bag is
seriously affected by the incoming melt temperature. So the heat

69

transfer eld should have been taken into account to correctly


predict the distortions of the bag and the corresponding velocities
in their modeling studies.
Fortier et al. [11] studied the effects of distributor bag stiffness in
the presence of a skim dam on the temperature and velocity elds
without taking into account the solidication aspect of the process.
They carried out both physical water and mathematical modeling
studies. The authors observed that in their water model a rigid
distributor bag directed the water downwards into the sump
whereas a regular soft combo bag directed the water upward and
towards the meniscus. By measuring the temperatures at the
meniscus during real casting experiments they obtained lower
meniscus temperatures for the rigid distributor bag than for the
soft bag, which they attributed to the downward melt outow into
the liquid sump for the rigid bag. The authors used commercial
ANSYS/CFX software to model the top part of the caster along with
its distribution system. Nothing was mentioned in their modeling
of the velocity eld whether they considered the ow as laminar or
turbulent. Moreover, the authors did not compare the predicted
results with similar results from others.
Using the continuous casting module of the commercial software named ProCAST (v2006.1), Arsenault et al. [12] predicted 3D coupled laminar uid ow and thermal elds along with solidication of a DC caster tted with a submersed nozzle inside a
Thermally Formed (TF) semi-rigid combo bag supplied by PYROTEK
INC. The authors were interested in the steady state results but
unfortunately could not reach a steady state solution with the
transient implementation of the CFD code. They claimed that their
predicted transient velocity prole inside the impervious skirt of
the TF combo bag was visually similar in nature with their water
modeling results. They didn't compare their predicted results with
any published experimental or industrial data. Further, nothing was
mentioned how the solidication including mushy zone of the
process was modeled.
A 3-D nite element commercial code was used by Ilinca et al.
[13] to carry out the modeling studies of uid ow and heat transfer
of DC casting of aluminum ingots where the hot-top mold was
tted with a Thermally Formed (TF) combo-bag delivery system.
The authors have shown that the inlet ow rate near the vicinity of
the combo bag was dominated by the forced convection and by
natural convection outside this region. In order to accommodate
the ow though the fabric cloth of the combo bag they articially
took into account the porous structure of the fabric by describing in
advance the pressure drop there as a non linear function of vertical
velocity. Using the transient approach of the code, the authors
could not reach a converged steady state solution. In addition, the
authors did not validate their code.
The objective of the present work is to carry out a comprehensive 3-D numerical study incorporating the turbulent melt
ows, heat transfer and solidication in a coupled manner for the
simulation of industrial scale DC casting of aluminum-alloy slabs
during the steady state phase for the combo bag melt-feeding
system. The bottom plate of the combo bag is tted with a
porous plate (foam) so that the momentum of the incoming melt
is reduced and it allows the melt to achieve uniformity in terms of
outlet velocity. Another purpose of the porous bottom plate is to
arrest the solid oxide inclusions those are generated in the
launder/trough assembly delivery system above the caster. In this
study, the latter objective has not been carried out since a
completely separate study is required to track the movements of
inclusions in the presence of turbulent melt ow in the clear uid
region and within the foam lter. The present study is an initial
step towards understanding the complex interactions of the
various parameters of the DCC process for a combo bag melt
feeding system.

70

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

2. Mathematical model
2.1. Model description
In this study, a rectangular rolling ingot having dimensions of
2500 mm  1730 mm  660 mm in the x, y, and z directions,
respectively, was considered as the physical model which is schematically represented in Fig. 1. A standard hot top mold having a
length of 210 mm was considered. The hot top consisted of an
insulated part of length 130 mm (not shown in Fig. 1) while the lower
part of 80 mm formed the active mold. Since the modeled domain
has a two-fold symmetry, hence only quarter (ash color region of
Fig. 1) of the domain was modeled to decrease the computational
time and associated costs. This gure also shows the origin and the
directions of the coordinate axes used. The melt was being delivered
to the mold through a newly proposed combo bag. The structure of
the combo bag is such that it formed a rigid channel which is closed
on the two rolling sides and the top while there are two openings
facing the two narrow sides of the caster. The bottom plate of the
combo bag is made of stainless steel porous foam so as to allow a
controlled delivery of the melt as well as to capture some of the solid
oxide inclusions that might had been generated in the launder/
trough assembly above the combo bag. Fig. 2 shows the enlarged
view of the newly designed combo-bag. A square-shaped nozzle
having a cross-sectional area of 900 mm2 supplies the melt into the
hot top from the launder/trough assembly. The nozzle walls, the
walls of the distributor bag, and the bottom plate were all assumed
to have a constant thickness of 5 mm [8] and were assumed to have
been made of stainless steel. The thermo-physical properties of
stainless steel are given in Table 1. The nozzle was immersed in the
liquid metal pool to a depth of 20 mm and placed at the center of the
distributor bag, and the whole nozzle-distribution bag assembly was
placed at the geometrical center of the caster. The dimensions of the

Fig. 2. Schematic of a combo bag having a bottom porous-foamed plate.

bag were 35 mm  400 mm  180 mm in the x, y, and z directions,


respectively.
2.2. Test material
The casting material was considered as an almost pure
aluminum alloy AA-1050 (99.5% aluminum plus 0.5% other alloying

Fig. 1. Schematic of a vertical DC caster with the calculation domain represented by ash color for a submerged nozzle with a combo-bag.

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

Launder and Sharma [16] is used. The governing equations and


turbulence model are well explained in Begum [17].

Table 1
Physical properties of the stainless steel [14].
Properties (symbol)

Value (units)

Thermal conductivity (solid) (ksolid)


Density (solid) (r)
Specic heat (solid) (CP)

0.036 (kW/m-K)
7850 (kg/m3)
0.690 (kJ/kg-K)

elements) which has a small solidication range of 11  C. The


physical properties of AA-1050 alloy were taken from Zuidema
et al. [15] and are summarized in Table 2.
2.3. Prescribed assumptions
In order to realistically model the problem, the following assumptions were considered in developing the current mathematical model:
[1] A xed-coordinate system (Eulerian approach) was
employed in the simulation.
[2] Local thermodynamic equilibrium during solidication was
assumed to prevail. Heat ow was taken to be very fast and
every point reached equilibrium with its neighboring points
instantaneously.
[3] A mushy-uid solidication model was assumed and pore
formation was ignored.
[4] Flow in the mushy region was modeled similar to a ow
through a porous medium.
[5] Molten aluminum was assumed to behave as an incompressible Newtonian uid and turbulence effects were
approximated using the popular two equation ke model of
turbulence.
[6] Evolution of latent heat in the solidication domain was not
inuenced by the microscale species transport.
[7] Variation of liquid fraction in the mushy zone was assumed
to be a linear function of temperature.
[8] The thermo-physical properties of aluminum were invariant
with respect to the temperature and there was no viscous
dissipation effect. The thermal buoyancy term was incorporated in the momentum equations by employing the Boussinesq approximation.
[9] The top free surface was assumed to be insulated.
[10] Only the steady-state situation was modeled without taking
into account the transient start-up condition.

2.4. Numerical formulation


2.4.1. Governing equations for turbulent melt velocity and enthalpy
The 3-D general governing transport equations for turbulent
melt velocity and enthalpy at steady state in Cartesian tensor notation can be written is as follows:

vrui F
v
vF
GF

vxi
vxi
vxi


SF

71

(1)

where, r is the melt density and ui is the velocity component in the


xi e direction and F can be the dependent variables such as 1 for
continuity equation, velocity ui for momentum equations, temperature T for energy equation, turbulent kinetic energy k and rate
of energy dissipation for k and equations. GF is the effective
diffusion coefcient of F and SF is the source term of F. The gravity
direction is in the positive x-direction. The low-Reynolds number
version of the popular ke eddy viscosity concept proposed by

2.4.2. Solidication modeling


The major barrier in analyzing phase change problems for alloys
(impure materials) is that two interfaces appear, namely, solidmushy and mushy-liquid interfaces whose locations are unknown
a priori, and across which the latent heat is released. At the phase
boundaries, continuity of velocity and shear stresses do prevail.
Also, the continuity of temperature is to be respected at each
interface. In addition, an account has to be made in the jump of the
heat uxes across each interface, which is proportional to the latent
heat. To avoid tracking of the unknown interfaces, a single domain
approach is followed in the present simulation. Among the various
xed domain techniques that take into account the release of latent
heat, are apparent or effective heat capacity method and enthalpy
based method. In the current work, the enthalpy based method,
originally proposed by Voller and Prakash [18] for laminar ow
solidication problems has been extended for the 3-D turbulent
solidication problem [17,19]. In this method, the total enthalpy is
decomposed into sensible and nodal latent heat in the energy
equation.

H h DH

(2)

where h is the sensible heat dened as:

Zh

ZT
dh

cp dT

(3)

Tref

href

For constant cp, and taking href 0 at Tref, the above equation can
be integrated to obtain the sensible heat as:



h cp T  Tref

(4)

In order to establish the region of phase change, the latent heat


contribution was specied as a dependent variable with respect to
the local temperature, i.e., DH f(T).
The energy equation was applied for the entire calculation
domain including solid, liquid, and mushy regions, by changing the
nodal latent heat as a specic function of liquid fraction, that is

DH DHf fl

(5)

where fl is the liquid fraction. It is assumed in the present model


that fs (1  fl) increases linearly from zero at the liquidus temperature to 1 at the solidus temperature. Mathematically the above
statement can be represented by the following expression in terms
of equivalent liquid fraction fl.

fl

8
>
>
>
>
<

whenT >Tliquidus 0Liquidregion

T Tsolidus
whenTliquidus T Tsolidus 0Mushyregion
>
T
liquidus Tsolidus
>
>
>
:
0
whenT <Tsolidus 0Solidregion
(6)

where Tliquidus or Tl and Tsolidus or Ts are the liquidus and solidus


temperatures, respectively. DH is the nodal latent heat of a specic
control volume which varies between 0 to DHf at Ts and Tl,
respectively.
It is to be noted here Tl and Ts are constant in the present model.
Ideally, for an alloy one needs to solve the micro-segregation
problem in order to allow changes of Tl and Ts as solidication

72

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

empirical measurement of the permeability of the porous media.


Darcy's law for a porous media can be written as:

Table 2
Physical properties of AA-1050.
Variable (units)

Value

Thermal conductivity (liquid or solid) (kW/m-K)


Specic heat (liquid or solid) (kJ/kg-K)
Latent heat of fusion (kJ/kg)
Liquidus temperature ( C)
Solidus temperature ( C)
Viscosity (kg/m-s)
Density (liquid or solid) (kg/m3)

0.231
0.90
396.4
657.0
646.0
16.2  104
2700

progresses. This would have enormously complicated the modeling


effort, particularly for three-dimensional modeling of mushy zone
solidication in the presence of turbulent melt ow. Hence, the
above simplication was adopted.
The nal dimensional energy equation can be written in the
following form:

"




 #
v rui H
v rui DH
v
m mt vh


vxi
Pr st vxi
vxi
vxi

(7)

The last term in Eq. (7) represents the latent heat exchange rate
due to solideliquid phase change resulting from the turbulent
convective ow. The above equation was suitably nondimensionalized and its nal form is listed in Table 3.
2.4.3. Modeling of uid ow in the mushy region
In this study, in order to solve the complex mushy uid problem,
it was assumed that the solidication is predominantly equiaxed
and the solid and liquid move at the same velocity in the mushy
region which is different from the casting speed and is not necessarily equal to the adjacent melt velocity. When the melt is
completely solidied, the shell moves at the casting speed. In reality, the transport phenomena that prevail in the mushy region are
quite complex and difcult to describe by macroscopic equations.
Because of the simplicity of the Darcy law for porous media, in the
current model this law is adopted. This law is based on the

Table 3
Summary of the non-dimensional governing equations for the melt region.
*

G*F

S*F

Equation

Continuity
U-Momentum

1
U

0
1
Re 1

m*t

V-Momentum

1
Re 1

m*t

1
Re 1

m*t

W-Momentum W
Kinetic energy k*



vU  A* U  U  Grm h*  h*
G*F vX
s
a
Re
i
Re2


*
* vV
v
A
vXi GF vXi  Re V  Vs


v
A* W  W
vX
G*F vW
 Re
s
vXi
i

v
vP
vX vXi
*

vP
vY

vP
vZ

D*k

G*

Re  Re Sk*
1 skt


E*
Rate of energy * 1
1
* *
*2
m*t
Re f1 C1 G k*  C2 f2 k* Re2 S*
Re 1 s
dissipation


*


vVDH *
vWDH *
 vUDH
Energy
h* 1 1 m*t
vX vY
vZ

st
Re Pr

vUi
vXj

!
vU

vUi
vXj ;

vXij
!

E* 2m*t

m*

(8)

where K0 is the permeability, which is a function of porosity, or in


the case of a mushy region of a binary alloy, a function of liquid
fraction. The permeability decreases with decreasing liquid fraction
and ultimately forces all the velocities to become zero in the case of
a stationary solid or in the case of vertical DC casting it forces the
solid to take the casting speed. Incorporating the diffusive and
convective momentum ux terms, one can write the modied
Darcy equation in the following dimensional form:

"
#



v rui uj
v
vui
m

ml mt
 SU  0 ui  uis
i
vxj
K
vxj
vxj

(9)

where the coefcient m/K0 decreases from a large value in the solid
phase to zero in the liquid phase. Consequently, the Darcy term (the
third term in the right hand side of Eq. (9)) vanishes as the liquid
fraction becomes one. It is to be noted that the Darcy term is an
adhoc sink term. The CarmaneKoseny equation is adopted for the
relation between the permeability and the liquid fraction.

m
C1  fl 2

K0
f3 q

(10)

where q is a small positive number introduced to avoid division by


zero during the numerical calculations. C is a constant that depends
on the morphology of the porous media. The value of C was estimated from the expression given by Minakawa et al. [20].

.
C 180 d2

(11)

where d was assumed constant and was equal to the secondary


dendrite arm spacing (DAS). The value of d is of the order of
1.0  104 m. A large C value forced the velocity to be equal to the
casting speed.
2.4.4. Non-dimensionalization of the governing equations and
boundary conditions
In order to obtain the relevant parameters and for the generalization of the results it is a better practice to solve the nondimensionalized from of the governing equations and associated
boundary conditions. The following dimensionless variables were
used to non-dimensionalize the governing partial differential
equations and boundary conditions:

x
y
z
u
v
w
; Y ; Z ; U
; V
; W
; P*
D
D
D
uin
uin
uin
P
k
D
h
DH
m
; k* 2 ; * 3 ; h*
; DH*
; m*t t
DHf
DHf
m
ro u2in
uin
uin
(12)

where.
Re rumin D :
G* m*t

1
Re



K 0 vP
ui 
 rgxi
m vxi

v2 U i
vXj vXk

p
*
D*k 2 vvXki

!
v2 Ui
vXj vXk

p
v k* :
vXi
3:4
*

*2

*2

; fm* e1Ret =50 ; Ret Re k* ; m*t ReCm fm* k* ; f1 1;


2

t
2

f2 1  0:3eRet ; A*

C * 1  f1 2
:
fl3  q

Cm 0.09, C1 1.44, C2 1.92, sk 1.0, s 1.3, st 0.9 q 1  10

30

where D is the hydraulic diameter of the nozzle, uin is the nozzle


inlet velocity and DHf is the latent heat of solidication. All the
conservation equations can be expressed in a general form of a nondimensional partial differential equation. The Cartesian-tensor
form of this equation is:





v rUi F*
v
vF*
G*F
S*F

vXi
vXi
vXi
.

i 1; 2; 3

(13)

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

The values of F*, and the associated denition of G*F and S*F , for
all of the transport equations are listed in Table 3. The development
of a general-purpose program is considerably simplied because of
the casting of the governing equations in a general form.
The non-dimensional form of the boundary conditions becomes:
[1] Nozzle inlet:

.
U 1; V W 0; h* h*in ; k* 0:01; * cm 0:013=2 0:05
(14)

[2] Free surface:

vV vW vk* v* vh*

0;
vX
vX
vX
vX
vX

U0

(15)

73

Air gap inside the mold :


.

g x 0:15 kW m2  K

200  x < 210 mm

(22)

Secondary cooling zone:


.

x  x1 
g x ggap
gmax  ggap kW m2  K
x2  x1


210  x < 220 mm

(23)


.

x  x2 
g
 gmax kW m2  K
g x gmax
x3  x2 film


220  x < 260 mm

(24)

g x gfilm

x  260 mm

(25)

where, x1 210 mm, x2 220, x3 260 mm and


[3] Symmetry planes:

At the XeY plane :

vU vV vk* v* vh*

0;
vZ
vZ
vZ
vZ
vZ

W0
(16)

At the XeZ plane :

vU vW vk* v* vh*

0;
vY
vY
vY
vY
vY

V 0
(17)

[4] Outlet:

vU vk* v* vh*

0
vX
vX
vX
vX

(18)

V W 0

(18a)

[5] Moving walls:

U Us us =uin ; V W k* * 0;


vh* vh*
gD  *
hs  h*a


K
vY
vZ
(19)

where g is the effective heat transfer coefcient between the solid


ingot surface and the surrounding, h*s represents the nondimensional enthalpy at the ingot surface and h*a represents the
non-dimensional ambient enthalpy. The values of the effective heat
transfer coefcient along the length of the slab were taken from
Vreeman and Incropera [21], which are given below:

Adiabatic section : gx 0: kW

.

m2  K 0  x < 130 mm
(20)

Mold  metal contact region :



.
130  x < 200 mm
gx 1:5 kW m2  K

(21)

.

.

ggap 0:15 kW m2  K ; gmax 20:0 kW m2  K ;
.

gflim 10:0 kW m2  K :
The air gap distance between the mold and the semi-solid shell
was assumed to be 0.4 mm which is within the range of
0.2e0.5 mm suggested by Prasad and Bainbridge [1]. It should be
noted here that in the present model the non-dimensionalized
transport equations and associated boundary conditions were
solved, which therefore generated non-dimensional results. To
make the results easily understandable, all predicted results in this
paper are reported and discussed in their primitive (dimensional)
forms.
2.4.5. Modeling of porous bottom plate of the combo bag
Fluid ow through a porous media is a complex phenomenon
due to the presence of solid matrix and the interactions of various
physical elds. To simplify the problem, the following assumptions
were also made with the assumptions listed in Section 2.3.
[1]. The lter was considered as non-deformable, homogeneous
and isotropic porous media, which was saturated with the
incoming melt.
[2]. The solid matrix of the porous media was considered to be in
local thermal equilibrium with the surrounding melt.
To model the DC casting process with internally placed porous
foam, two different sets of transport equations (continuity, momentum and energy) for clear uid and porous regions are needed.
For clear uid region the time-averaged turbulent form of the
governing conservation equations are already listed in Table 3. The
conservation equations for the porous region were based on
BrinkmaneForchheimer extended Darcy model. Beckermann et al.
[22,23] have discussed the importance of the two extensions of the
Darcy model. These authors have found that for modeling porous
media with a high permeability (i.e., a high Darcy number) the
aforementioned two extensions are necessary. It should be
mentioned here, the linear additions of the above two terms in the
Darcy model are based on semi-empirical models from the experimental ttings of the results. In this study, the ow of the melt
inside the porous lter was considered to be turbulent and
incompressible. The mathematical description of the turbulent ow
in the porous region is still a controversial issue. The present
mathematical model followed the approach where a volume-

74

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

averaged operator was applied to the local time-averaged turbulent


equations, which was proposed by Pedras and de Lemos [24e27]
for turbulent ow in porous medium.
In terms of the supercial (Darcian) velocity, the volumeaveraged conservative forms of the transport equations for turbulent ow in Cartesian-tensor notation for the porous layer are
[28e31] as follows:

Continuity equation :

 
1 vui
*
0
f
vxi

(26)


 
v rui uj
1
*
vxj
f2
"
!
#

 
CF r uj ui
vP
1
v
vui vuj
m
0
0
*


f*ml
 rui uj  ui  p
vxi
K
vxj vxi
f2 vxj
K


Momentum equation :

rgx
(27)

Energy equation :
v

vxj


  
1 v ruj H
*
f
vxj

vT
 f1=fg*ru0j H0
kEffective
vxj

!
(28)

Transport equation for turbulence kinetic energy:

"
!

  
 #  . 
1 v ruj k
v
mt vk
vui vuj vui
2
*

ml
1 f *mt
f
vxj
sk vxj
vxj vxi vxj
vxj

rk uj
Ck p  r  Dk
K
(29)
Transport equation for rate of dissipation of kinetic energy:

"


 #
v ruj
v
mt v

1=f *

ml
c1 f1 G  c2 r
vxj
k
s vxj
vxj



u j
2
rE
 Ck p f 2
k
K


(30)

where, K permeability of the porous media; f is the porosity


(void fraction) of the porous matrix; CF Forchheimer coefcient.
As in Vafai et al. [32], CF can be represented as:
q
CF 1:75= 1501=f3 ; the value 0.55 of CF was used in this work.
keffective effective thermal conductivity fkuid (1  f)ksolid in
the energy equation. The value of Ck was taken as 0.28 as suggested
by Pedras and de Lemos [26].
Here, u represents the time-averaged velocity vector;
0
u turbulent uctuations; k turbulence energy 1=2u0i u0j ;
turbulence dissipation rate; ru0i u0j Reynolds stresses in the
momentum equations; and ru0j H 0 the turbulent heat ux in the
energy equation. The right hand side of the momentum equation
contains the pressure gradient term, the viscous diffusion term
(Brinkman term), the turbulent diffusion term, the Darcy term, the
Forchheimer term and the gravitational term.
The macroscopic Reynolds stress ru0i u0j were also modeled by
using
0

the

eddy

viscosity

concept

and

was

given

by:

rui uj mt vui =vxj vuj =vxi 2=3dij k; and the turbulent viscosity

was given by: mt rcmfmk2/.


In the above Eqs. (29) and (30), sk, s, c1, c2, cm dimensionless
constants; and f2 and fm damping functions. The coefcient f1 is a

function that increases the magnitude of the destruction of the


kinetic energy of turbulence () near the wall. Launder and Sharma
[16] low-Reynolds model was used here to take into account the
low turbulent Reynolds number effects. The following damping
functions f2 and fm suggested by Launder and Sharma were used:

2
6
6
fu exp6
4

3
7
3:4
7
2 7

5
1 RT=50



f2 1  0:3 exp R2T

(31)

(32)

where RT k2/n turbulent Reynolds number; and n is the kinematic viscosity. Following Launder and Sharma (1974), the model
constants used here are given by: Cm 0.09, C1 1.44, C2 1.92,
sk 1.0, s 1.3, and the value of f1 is set to 1.
Similar to the transport equations for the melt, the porous media
transport equations (Eqs. (26)e(30)) were non-dimensionalized
with the same dimensional variables and were cast in the general
form as Eq. (13) with appropriate coefcients, and source and sink
terms. It is to be noted here that the non-dimensional transport
equations for the porous bottom plate of the combo-bag contained
two additional parameters, porosity(f) of the medium and Darcy
number (Da K/D2) both of which needed to be prescribed before
solving the equations.
The control volume based nite-difference scheme was used to
solve the modeled equations and boundary conditions which
offered a great advantage by the fact that the discretized governing
equations were derived by integrating the partial differential
equations over each control volume of the calculation domain
which automatically handled the strong momentum and heat exchanges at the interface of the clear uid and porous lter when the
relevant thermo-physical properties of the porous plate were
appropriately prescribed. Therefore, the boundary conditions at the
porous-uid interface were not necessary to consider.
2.4.6. Solution procedure
A control volume (CV) based nite difference approach was
adopted to discretize the equations. Staggered control volumes
were employed in the computational domain. A hybrid difference
scheme [33] was used to discretize the convectionediffusion terms.
In the present study there are seven variables, namely, U, V, W, P*,
k*, *, h* which were solved sequentially to obtain a converged
solution [33]. For each of the transport variable, the discretized
equation was cast in the following general form:

ap fp aE fE aW fW aN fN aS fS aT fT aB fB bf

(33)

where aE, aW, aN, aS, aT, and aB are the coefcients of the six
neighboring nodes of node P of a box-shaped rectangular control
volume. The term bF represents the linearized source term. The
convection and diffusion terms are embedded in the coefcients,
depending on the transport variable, the source term may contain
various terms such as pressure gradient, buoyancy force, and Darcy
terms. To resolve the velocityepressure coupling in the three momentum equations, the SIMPLE algorithm [34,35] was used. In
actual implementation of the SIMPLE algorithm, an iterative solution scheme was adopted using an implicit relaxation technique to
solve the discretized equations. The following combination of
under-relaxation factors provided converged solutions:

aU 0:1; aV 0:1; aW 0:1; ap* 0:1; ah* 0:1;


ak* 0:1; a* 0:1; and am* 0:3

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

The discretized equations for each variable were declared to


have converged when the sum of the residuals for that particular
variable (RF) was less than 0.001. The convergence criterion
described above can be dened mathematically as follows:

RF




X
X

anb Fnb  b
ap Fp 

all nodes

(34)

nb

The CPU time per iteration was about 1.0 min. The computations
were performed on a personal computer having a speed of
2.66 GHz and tted with an RAM of 4 Gigabytes. For a typical case,
to obtain a fully converged solution it took about 25 days and
required 36,000 iterations. For various parametric studies, in order
to decrease the computational time, the runs were restarted from
the converged solutions of the variables obtained for different
parametric cases.
3. Grid independency test
The grid independency tests were performed using three sets of
grid systems, namely, 60  42  24, 70  52  34, and 80  62  44
(x, y, z directions), which constituted 51,040, 108,800, and 196,560
control volumes, respectively. The local surface heat uxes at
y 430.5 mm, and z 322.5 mm for x varying from 135 to 400 mm
from the top free surface were compared for the above three grid
arrangements. It was found that the local surface heat ux varied by
less than 5% between the coarse and ne grid systems. The details
of the predicted results along with the comparisons are provided in
Begum [17]. In order to conserve space, those results are not reported here. To optimize CPU resources with an acceptable level of
accuracy and reasonable computational time, all production runs
reported in this paper were performed with a grid distribution of
60  42  24 grid points (or 51,040 control volumes).
4. Validation of the present developed code
Since a comprehensive 3-D DC casting work for a slab of
aluminum alloy AA-1050 similar to the one simulated in this study
is not available in the literature, hence the thermo-physical
properties of AA-3104 were considered and modeled the vertical
DC casting for the latter alloy in order to compare with the
experimental measurements of Jones et al. [36]. The latter authors
provided the measured temperatures at the ingot wide symmetry
plane for a rolling ingot of AA-3104 alloy. The simulated ingot had
a dimension of 2500  1320  660 mm in the x, y and z directions,
respectively and was identical to that used in the real casting
experiment by the above authors. The positions of the solidication front from the experimental measurements and predicted
from the present numerical model are compared in Table 4. It is
apparent from the above comparison of the depth of the solidication front that closer to the slab center, the predicted and
measured distances matched quite well, while further the thermocouples were away from the center and near the side wall the
match is not that satisfactory. This could be due to the fact that
Jones et al. did not measure the thermal boundary conditions
experimentally, and as a result, the employed thermal boundary
conditions used in the model might not have been appropriate.
Closer to the ingot center, a good agreement obtained with regard
to the vertical distances of the solidication front is an indication
that the heat transfer, namely, conductive, convective, and solidication which are at play in the caster as well as the turbulent
aspects of the process are well incorporated in the mathematical
model. Further details about the verications of the code are
available in Begum [17].

75

Table 4
Comparison of experimentally measured and numerically predicted vertical distance of the solidication front from the top of the mold at the wide symmetry plane
(z 0) for aluminum (AA-3104) alloy.
Jones et al.
Number
Horizontal distance
of sacricial
from the narrow face expts. [36]
(mm)
thermocouple of the mold wall
at the wide symmetry
plane (mm)

Relative
Present
numerical percent error
prediction
(mm)

1
2
3
4

140
281
382
510

45
147
249
351

63.8
191.3
418.2
568.6

119.4%
46.89%
8.66%
10.31%

5. Results and discussion


To facilitate the discussion a summary of the relevant parameters and their values are given in Table 5. For cases (1e4) of Table 5,
the velocity and temperature elds will be discussed rst which
will be followed by a discussion of other relevant quantities,
namely, sump depth and thickness of the mushy layer developed at
the ingot center. The latter discussion will be followed by the
growth of the solid-shell at the exit of the mold and the surface
temperatures at four strategic locations along the cast. Since the
results for cases (5e8) of Table 5 are qualitatively similar to the
earlier cases, therefore only the solid-shell thickness for various
effective metal-mold contact heat transfer coefcients (HTCs) is
presented and discussed. It is to be recognized here that for a xed
casting speed, the melt inlet velocity at the nozzle was adjusted by
satisfying the overall mass balance. Since the ratio of the crosssectional area of the nozzle to caster is approximately 1:1270,
hence the velocity of the melt at the nozzle inlet is 1270 times
higher than the casting speed. As a result the ow of the melt in the
hot top is turbulent even for the lowest casting speed of 40 mm/
min. Another point to note is that in the present model, to incorporate the thermal buoyancy effect, the modied Grashof number
(Grm) was considered to be 109, which was kept xed for all cases.
Since the effect of melt superheat on the solidication prole was
found not to be very signicant (Begum [17]), hence this parameter
was kept constant at 32  C. On the basis of the standard porous
foam structure, which are usually used for the ltration of
aluminum alloys, a porosity (f) of 0.9 [37] and permeability (K) of
9.0  108 m2 [38], respectively were taken and these values were
not varied in this work. For this geometry the permeability corresponds to a Darcy number (Da) of 1.0  104. The parameter Da (K/
D2) appeared in the non-dimensionalized momentum equations
and its value had to be prescribed in advance.
5.1. Velocity eld with superimposed streamlines and temperature
distributions
Fig. 3(i)e(iv) shows 3-D surface plots of the predicted temperature contours and velocity elds with superimposed streamlines
for four cases (1e4). Each gure shows two symmetry planes [wide
(xey plane) and narrow (xez plane)] as well as the top ingot surface
of the calculation domain. Fig. 3(b, d, f, h) shows the corresponding
3-D velocity vector elds with streamlines. From these gures one
can see that there are two streams of hot melt coming out from the
distributor bag at each casting speed. The stream which is emerging
from the side window of the combo bag referred as the horizontal
jet, and the other stream coming out through the bottom porous
plate of the distributor bag is referred to as the downward vertical
jet. The strength of the horizontal jet is reduced because a portion
of the incoming melt from the nozzle is exiting from the bottom
porous plate. It is to be recognized here that the porous plate is very

76

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

Table 5
Summary of the various parametric conditions studied for the combo bag with a metallic foam bottom plate.
Case

Inlet melt
superheat, [ C]

Casting speed,
us, [mm/min]

Metal-mold contact,
HTC [W/m2-K]

Inlet Reynolds number, (Re)

Inlet Peclet number,


(Pe Re  Pr)

1
2
3
4
5
6
7
8

32
32
32
32
32
32
32
32

40
60
80
100
40
40
80
80

1500
1500
1500
1500
750
3000
750
3000

42,367
63,550
84,734
105,918
42,367
42,367
84,734
84,734

267
401
535
669
267
267
535
535

Fig. 3. 3-D surface plots of the temperature contours and velocity elds with streamlines for complete solution domain for cases: (i) case-1 (a, b); (ii) case-2 (c, d); (iii) case-3 (e, f);
(iv) case-4 (g, h).

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

active just beneath the nozzle area where the jet momentum is
high. This is reected by the vertical arrows reaching a greater
distance in the axial (x) direction in the nozzle zone of the combo
bag. The horizontal jet of hot melt, after reaching the narrow side of
the slab, is seen to split into two streams. One stream is forming a
counter clockwise vortex at the wide symmetry plane, hence forth,
will be called the lower recirculation zone. Another part of the
steam moves upward and then it travels along the narrow and wide
faces and nally reaches at the mid-plane of the wide face where it
divides into two streams. One stream forms a clockwise recirculation zone at the top free surface, which is referred to as upper
recirculation zone, while the other stream moves in an inclined
manner, termed as angle ow, and meets with the ow of the lower
recirculation zone. In order to depict clearly the above description
of the ow eld, an enlarged view of the velocity vectors with
streamlines are portrayed in Fig. 4(a, b) for low and high casting
speeds of 40 and 100 mm/min, respectively for only the top part of
the caster. Since the velocity vectors inside the nozzle and combo
bag are very large compared to the velocity eld outside the above
two regions, hence while plotting the ow elds the velocity vectors inside the nozzle and combo bag were set to zero to overcome
the scaling problem.
One objective of the melt distributor in DCC is to minimize
turbulence at the top free surface so that the ow does not promote
the generation of the oxide inclusions there. Another objective of
having reduced ow there is to prevent the carryover of the inclusions into the cast. The proposed melt feeding system shows
that an upper recirculation zone forms at the free surface which
may not be desirable for a DCC process. A comparison of Fig. 3(b, d,
f, h), clearly shows that with the increase of the casting speed the
strength of the ow is increased. This can be observed by
comparing the indicated maximum resultant velocities reported in
the above four gures. This is mainly due to the additional
incoming melt for the higher casting speeds.
Fig. 3(a, c, e, g) shows the temperature contours along with the
solidication front (solidus) and the liquidus isotherm. The effect of
casting speed on temperature elds can be seen by investigating

77

the above gures. Isotherm plots in these gures indicate that


increasing the casting speed is increasing the liquid melt region
inside the domain which is represented by the location of the liquidus isotherm. The mushy region is conned by the solidus
(646  C) and liquidus (657  C) temperatures of the alloy. The solid
part of the ingot, according to the model assumptions, is traveling
axially downward with the casting speed. The same gures reveal
that the solidication front is becoming steeper and moving
downward with the increase of the casting speed. Compared to the
mushy zone, one can see in each of the above gures that the liquidus isotherm is completely non-uniform and forms a cliff at
around the middle of the wide symmetry plane (of the computational domain) and close to the side window of the combo-bag. This
has caused an increase in the thickness of the mushy layer there
compared to the other regions of the cast. This marked-lift of the
liquidus isotherm is due to enhanced thermal convection currents
resulting from the merging of the angle ow and the lower recirculation zone. With the increase of the casting speed this cliffshaped mushy region is moving downward and is taking different
shapes due to the stronger thermal convection there. A similar cliffshaped region was reported in a recent paper by Zhang et al. [39] in
their 2-D modeling of an AA-7050 round billet caster for a vertical
melt feeding system through a delivery nozzle.
Figs. 5(a, b, c)e8(a, b, c) illustrate longitudinal 2-D view of
temperature and velocity proles at the wide symmetry plane
(z 0) and parallel to the wide symmetry plane at z 62.5 mm, and
z 312.5 mm for the aforementioned four cases. In presenting the
results, only about two-third portion of the ingot is taken and
magnied for the purpose of clarifying the results. The left hand
panels of these gures show temperature contours while the right
hand panels of these gures illustrate the velocity vectors at three
longitudinal cross-sectional planes. The lower recirculation vortex
at the wide symmetry plane is clearly visible in the 2-D velocity
vector plots of Figs. 5(a)e8(a), and its strength is increasing with
increasing casting speed. The strength of the ow is represented by
the magnitude of the maximum resultant velocity. The trend of the
melt ow at z 62.5 mm and z 312.5 mm for four casting speeds

Fig. 4. 3-D enlarged view of the velocity elds with superimposed streamlines for the top part of the calculation domain for cases: (a) case-1; (b) case-4. The maximum resultant
velocity developed in the ingot outside the combo bag and the nozzle has also been indicated by the red color arrow above each gure. (For interpretation of the references to colour
in this gure legend, the reader is referred to the web version of this article.)

78

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

Fig. 5. Enlarged 2-D view of temperature contours and velocity vectors of the top domain for case-1 at: (a) wide symmetry plane at z 0 mm, (b) vertical plane parallel to the wide
face at z 62.5 mm, (c) vertical plane parallel to the wide face at z 312.5 mm.

are presented in Figs. 5(b)e8(b), and Figs. 5(c)e8(c), respectively.


As one moves from the wide symmetry plane towards the rolling
face one can see that, in the longitudinal plane at z 62.5 mm, the
strength of the lower recirculation zone has signicantly reduced. It
is for the rst time that the current mathematical model results
show the critical role played by the downward stream of melt issued from the bottom foam-plated combo bag in shaping the solidication front prole which will be explained later.
From Figs. 5(a)e8(a), one can see the solidication prole is
more horizontally-uniform along the wide and narrow symmetry
planes at the slab center, which may be, due to the enhanced
conductive mode of heat transfer in those locations. For the casting
speed of 40 mm/min, the horizontally-at part of the solidication
front is approximately 58 mm, measured from the center towards
the narrow slab face, whereas, this horizontal distance for casting
speeds of 60, 80, and 100 mm/min is increased to approximately

140, 152, and 165 mm, respectively, from the slab center, as illustrated in those gures. The increased uniformity of the solidication front with the increase in the casting speed is a very desirable
metallurgical characteristic in DC casting for the following reasons.
First, increasing the casting speed implies an increase in the
production rate of aluminum slabs. Second, an increase in the at
part of the solidication front implies that uniform heat transfer is
taking place over the increased surface area of the solidication
front. The latter would allow the slab to have better homogeneous
metallurgical characteristics. For a lower casting speed, the reason
for getting an almost-horizontally-attened solidication front is
attributed mainly due to the inability of the weak vertical jet
coming from the bottom porous plate to penetrate the stronger
angle ow near the central region, thereby the thermal convection
effect is reduced in that region. An increase in the casting speed
causes the speed of the vertical jet to increase. Consequently, a

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

79

Fig. 6. Enlarged 2-D view of temperature contours and velocity vectors of the top domain for case-2 at: (a) wide symmetry plane at z 0 mm, (b) vertical plane parallel to the wide
face at z 62.5 mm, (c) vertical plane parallel to the wide face at z 312.5 mm.

collision is taking place near the central region between the vertical
jet and the angle ow, thus reducing the strength of the resultant
ow and hence the thermal convective effect there. Moreover, for a
higher casting speed, the melt temperature is higher near the
bottom of the porous plate compared to the temperature of the
angle ow, as a result, the melt is becoming thermally stratied
there and causing less thermal buoyancy effect.
The left hand panels of Figs. 5(b)e8(b) show temperature
contours for a plane parallel to the wide symmetry plane at
z 62.5 mm. The temperature isotherms in each panel are seen
to have lifted upward in comparison to the isotherms presented
in the wide symmetry plane. The upward movements of these
isotherms, is mainly due to the weak convection within the
melt itself, and also due to the comparatively nearer location
with respect to the wide face where greater heat transfer takes
place.
Similar to the earlier gures, the left hand panels of Figs. 5(c)e
8(c) show temperature contours for a plane parallel to the wide

symmetry plane at z 312.5 mm, which is close to the wide face.


The velocity eld shows the melt is moving horizontally from the
narrow face toward the slab center in the hot top region and the
magnitude of the maximum resultant velocity is much lower
compared to the corresponding values in other two longitudinal
planes. From the temperature contours it is evident that most of the
melt is solidied there, with only a narrow and uniform horizontal
mushy region is left to be solidied. Since these panels are very
close to the wide face, as a result, the solidication is almost
complete there. A comparison of the isotherms among the four
panels shows that with the increase of the casting speed isotherms
move rapidly downward.
Fig. 9(a)e(d) gives numerical results for transverse crosssections (yez planes), parallel to the slab-free surface, for four
cases (1e4) of Table 5. These gures show the liquid, mushy, and
solid regions across the ingot cross-sections. Here, only the parts
showing the extent of the sump are presented and magnied. To
convey the results, seven transverse cross-sections of the ingot are

80

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

Fig. 7. Enlarged 2-D view of temperature contours and velocity vectors of the top domain for case-3 at: (a) wide symmetry plane at z 0 mm, (b) vertical plane parallel to the wide
face at z 62.5 mm, (c) vertical plane parallel to the wide face at z 312.5 mm.

taken, which are at an axial distance of 100, 200, 300, 400, 500, 600,
and 700 mm from the top ingot surface. In these gures, the rst
slice is taken near the end of the adiabatic section of the hot top at
an axial distance of x 100 mm. The second cross-section is taken
at the end of the metal-mold contact region within the mold, at
x 200 mm. The latter ve sections are located in the secondary
cooling region. Examinations of the gures show how the solidication proceeds along the wide and narrow faces of the ingot. At
x 100 mm, only liquid melt exist, as shown in these gures. Near
the end of the mold, a thicker solidied layer is formed along the
ingot rolling face compared to the ingot narrow face. The slow
moving melt along the rolling section allows more heat to be
extracted by the primary coolant of the mold. Whereas, the high
momentum melt stream delivered to the narrow side of the hot-top
mold through the side opening of the combo bag has delayed the
formation of the solidied layer in that part. The mushy layer is
separated between the center of the ingot and near the narrow slab

face, and becomes thicker there. From previous discussion on


various longitudinal cross-sectional views, it is evident that a
counter-clockwise vortex is formed right after the channel opening.
The angle ow, coming from the narrow symmetry plane, is
colliding with the vertical jet exiting from the bottom porous plate
and the resultant stream is moving upward and is merging with the
upwardly directed ow of the lower recirculation zone and is
leading to a wider thermal transition (mushy) region. At the
merging zone, these two streams restrict the advancement of the
liquidus isotherm, which is reected on the mushy layer formation.
Because of intense cooling in the secondary cooling region, the
growth of solid layer becomes greater which is demonstrated in
these gures. Considering the effect of casting speed, one can nd
that the solidication growth rate along the wide and narrow faces
is decreased due to the increase in the casting speed. With
increasing casting speed, the mushy layer is seen to exist at a higher
depth.

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

81

Fig. 8. Enlarged 2-D view of temperature contours and velocity vectors of the top domain for case-4 at: (a) wide symmetry plane at z 0 mm, (b) vertical plane parallel to the wide
face at z 62.5 mm, (c) vertical plane parallel to the wide face at z 312.5 mm.

5.2. Quantitative analysis for cases (1e4)


5.2.1. Predicted sump depth and mushy thickness at ingot center
Table 6 provides the predicted sump depth (vertical distance
from top of the mold to the solidication front at the center of the
ingot) and mushy layer thickness at the ingot center for four cases
(1e4). The sump depth and the mushy layer thickness are both
observed to increase with the increase in the casting speed, as
expected. The additional melt ow with the increase in the casting
speed leads to a downward shift of the solidus isotherm in the cast
direction at the ingot center.
From Table 6 one observes that the sump depth is approximately
466.715 mm for a drop rate of 40 mm/min. In comparison to the
lowest casting speed of 40 mm/min, for higher casting speeds of 60,
80, and 100 mm/min the relative difference in sump depth is about
4.75%, 5.93%, and 7.5% higher. At the ingot center, for a casting speed
of 40 mm/min the mushy thickness is about 7.01 mm, whereas, for
casting speeds of 60, 80, and 100 mm/min, the thickness increases

disproportionately by about 65.19%, 83.02%, and 124.54%, respectively compared to the casting speed of 40 mm/min. Upon comparisons of sump depth and mushy layer thickness for various
casting speeds one nds that the increase of the sump depth is not
that signicant with the change in casting speed while there is a
signicant increase in the mushy layer thickness with the casting
speed. Because of the collision between the heated vertical jet and
the relatively colder angle ow in the central area, the thermal
convection effect is markedly reduced. As a result, the effect of the
casting speed is not greatly manifested on the sump depth.
5.2.2. Predicted solid-shell thickness at two locations
In Fig. 10, the predicted shell thickness from the center of the
narrow face at the wide symmetry plane and from the center of the
wide face at the narrow symmetry plane at an axial distance of
210 mm from the top free surface is given in the form of a bar chart
for four cases (1e4) of Table 5. The predicted values of the shell
thickness in mm are also provided at the top inside each bar of the

82

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

Fig. 9. Contours of solidus and liquidus temperatures at various transverse cross-sectional planes (yez planes) of the top part of the ingot for four cases: (a) case-1; (b) case-2; (c)
case-3; (d) case-4.

chart for easy comprehension. For both locations, with the increase
of the casting speed the shell thickness decreases, as expected. For a
change of casting speed from 40 to 60 mm/min, the shell thickness
is seen to decrease rapidly whereas for the casting speeds above
60 mm/min the trend of the retardation of the solid-shell is
comparatively slow and gradual. This disproportionate change of
the solid-shell formation with the change in casting speed may be
due to the complex ow patterns produced by this melt feeding
scheme.
A signicant variation of the thickness of the solid-shell along
the narrow and wide faces at the exit of the mold is found. At the
mid-point of the narrow face, a thinner shell is seen to have
developed compared to the mid-point of the rolling face, and this is
true for all four simulated casting speeds. The reason behind this
phenomenon is the fact that the part of the melt coming out from

the side window of the combo bag travels in clockwise direction


along the narrow and wide faces before it splits into two streams
near the narrow symmetry plane. During this journey the melt
stream losses more momentum and heat, which has resulted in a
thicker shell around the wide face.
5.2.3. Predicted temperatures along the caster at four strategic
locations
It is to be realized that the surface temperatures of the cast in a
DCC process are usually not known. For quality control purposes, all
DCC operators would like to know the surface temperatures of the
cast as it comes out from the mold. Since the ingot surface temperature depends on the complex interactions of the cooling conditions in the mold as well as on a number of physical and
geometrical parameters of the process, hence the surface

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87


Table 6
Sump depth and mushy thickness at the ingot center for four cases (1e4).
Quantity

Sump depth (mm)


Mushy thickness (mm)

Values of casting speed (us) in mm/min


40

60

80

100

466.715
7.01

488.897
11.58

494.413
12.83

501.728
15.74

temperatures cannot be known in advance unless temperature


measurements are simultaneously recorded during the cast or are
predicted through veried numerical codes. An extensive search of
the DCC literature has revealed very little information on the rate of
heat extraction along the cast for different alloys operating under
various casting conditions. If the surface temperatures could be
known in advance one then could presumably determine where
the improvements in terms of heat extraction rates are needed to
obtain a good quality cast. With the above objectives in view, the
ingot surface temperatures are predicted for various imposed
effective heat transfer coefcients. The quantitative values of the
temperatures may further provide useful guidance in locating
possible cracking zones.
For the casting speeds of 40 and 100 mm/min, the temperature
distributions along the caster are presented in Fig. 11(a) and (b),
respectively, which correspond to cases (1, 4) of Table 5. Four
strategic locations are selected to portray the temperature distributions along the cast direction. These locations are: (a) center; (b)
mid-point of the wide face; (c) mid-point of the narrow face; and
the (d) corner point of the caster. As the effective heat transfer
coefcients (refer to Section 2.4.4) are increased from the mold to
the sub-mold regions, the increasing rate of heat extractions decreases the temperatures signicantly at all the four locations.
Further, both the gures show that from the top of the mold to the
impingement point the temperature drops very rapidly and after
the impingement zone the drop in temperature is rather gradual
except for the temperatures at the ingot center. Within the mold
and for part of the sub-mold region at the ingot center the temperature drops rather gradually and at a much slower rate
compared to the lm boiling zones in the lower part of sub-mold
region. In the latter region the temperature drops comparatively
at a higher rate until the end of the cast.
Up to a certain region of the secondary cooling zone, the midpoint of the wide face shows a slightly lower temperature
compared to the mid-point of the narrow face. Part of the heated

Fig. 10. Solid-shell thickness from narrow and wide slab faces at the wide and narrow
symmetric planes respectively at an axial distance of x 210 mm from the top free
surface versus casting speed for cases (1e4).

83

melt coming out through the side window of the bag is directed
towards the narrow face and as a result the surface temperature is
higher at that location. On the contrary, near the mid-point of the
rolling face, a relatively colder and slow moving melt generates a
lower thermal convective ow and has resulted in the reduction of
the surface temperature there. As a consequence of the two
different thermal convective effects in the said two regions there is
approximately 1.21e1.53 times thicker solid-shell is seen to
develop at the mid-point of the wide face compared to the midpoint of the narrow face (refer to Fig. 10) at the mold exit. This
trend is observed up to an axial length of 335 mm for the low
casting speed of 40 mm/min while for the higher casting speed of
100 mm/min this trend prevails up to a length of approximately
495 mm. Beyond the above axial distances, the trend in the temperature distributions is seen to have been reversed. In other
words, a comparatively higher value of the surface temperature is
predicted along the cast at the wide face compared to the value
predicted for the same axial location but at the narrow face. The
reasons behind this phenomenon are due to the higher thermal
resistance from the thicker solid-shell around the wide face and
also for the proximity of the mid-point of the narrow face to the
corner region where higher rate of heat extraction is taking place
from the two faces of the cast surfaces.

Fig. 11. Variations of surface temperature along the axial direction of the strand at four
locations of the caster for: (a) us 40 mm/min (case-1) (b) us 100 mm/min (case-4).

84

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

The surface temperature at the corner drops more sharply and


attains lower values compared to the mid faces. The reason behind
for this is stated earlier. This overcooled situation of the corner
compared to the mid-face regions and a little-bit overheated situation of the narrow face compared to the wide face have resulted in
the unequal increase in temperatures across the ingot crosssection. That may lead to the formation of surface and shape
defects.
A comparison of the two gures clearly further show that for a
higher casting speed the surface remains overheated for the same
axial location compared to the lower casting speed. This trend is
visible in all of the four ingot locations. The reason behind this is the
fact that the predictions were carried out for two casting speeds by
invoking the same convective heat transfer boundary conditions
without considering the enhanced cooling requirements for a

higher casting speed. As a result, at the water impingement point


for a casting speed of 40 mm/min the ingot surface temperatures at
the middle of the wide and narrow faces are seen to reach
approximately 190  C and 227  C, respectively, whereas for a
casting speed of 100 mm/min it reaches to a value of around 239  C
and 261  C, respectively at those locations. The surface temperatures for both cases show that the temperatures there are not
within the nucleate boiling regime (which is below 150  C) and
hence there could be the possibility of hot tears in those regions.
The cooling objectives at the steady state operations of all DCC
operators are to maintain the nucleate boiling of water near the exit
of the mold and down to the water streaming region. The nucleate
boiling is known to occur for a range of surface temperatures between 100 and 150  C [40]. The reason for trying to achieve
nucleate boiling is the fact that the highest heat extraction rate

Fig. 12. 3-D surface contours plots of non-dimensional turbulent viscosity of the top part of the ingot for four cases: (a) case-1; (b) case-2; (c) case-3; (d) case-4.

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

takes place from the slab surface to the cooling water under this
condition.
5.2.4. Non-dimensional eddy viscosity
The non-dimensional eddy/turbulent viscosity contours on the
top free surface and two vertically symmetric planes for four cases
(1e4) are shown in Fig. 12(a)e(d). Turbulent viscosity is a ow
property which represents the enhancement of molecular viscosity
of the melt due to the turbulent motion of melt there. In the present
study, the eddy viscosity model was used to calculate quantitatively
the level of turbulence. For this delivery system the turbulent viscosity varies from zero to approximately 1000 times the molecular
viscosity. A zero value of the turbulent viscosity appears in the solid
regions, and it increases with the increase of convective intensity
into the bulk liquid phase. The highest value of the turbulent viscosity is obtained at the exit of the side window of the combo bag,
as expected. From these gures, the higher value of turbulent viscosity is observed at the ingot top surface which is caused by the
upper recirculation vortex produced due to the studied melt
feeding scheme. This may contaminate the melt by carrying the
solid oxide scales and inclusions generated at the top free surface
into the cast. The above situation can seriously deteriorate the
quality of the as-cast products. A comparison among the four gures indicates that the level of turbulence increases with the
increasing casting speed which is consistent with the progressively
increased Reynolds number of the ow presented in Table 5 for
cases (1e4).
5.3. Inuence of effective heat transfer coefcient (HTC) on
solidication process
For vertical DCC processes in the mold region a higher value of
HTC ranging from 1000 to 5000 W/(m2-K) and a lower value
ranging from 100 to 500 W/(m2-K) have been reported in the
literature [3,41]. It is to be noted that not many experimental
studies concerning HTCs in the mold region for industrial-sized
slab casters of different alloys are available in the literature. In
this section, a sensitivity analysis has been carried out by changing
the value of HTC in the mold-metal contact region only. For this
purpose, three different arbitrary values of constant HTC for the
mold-metal contact region, ranging from 750 to 3000 W/(m2-K),
have been considered and the shell-thickness as a function of HTC
is plotted for two casting speeds, 40 and 80 mm/min as shown in

Fig. 13. Solid-shell thickness from the narrow slab wall at the wide symmetry plane at
an axial distance of x 210 mm from the top free surface versus effective heat transfer
coefcient [W/(m2-K)] for casting speeds of 40 and 80 mm/min for six cases (1, 3, and
5e8).

85

Fig. 13. The shell thickness is estimated at an axial distance of


210 mm from the top free surface at the wide symmetry plane from
the narrow face. It can be seen that the increase in HTC increases
the magnitude of shell thickness. Similar trends are found for both
casting speeds. For a xed HTC, the shell thickness decreases with
casting speed as noticed earlier. For a higher casting speed of
80 mm/min, an increase of the HTC from 1500 to 3000 W/(m2-K)
causes a signicant enhancement of solid-shell thickness, while a
relatively insignicant increase in the values of shell thickness is
predicted as HTC increased from 750 to 1500 W/(m2-K). For the
lower casting speed of 40 mm/min, less than 6% variation in the
solid-shell thickness is found due to the aforementioned change in
HTC. From this analysis one can conclude that for the lower casting
speed the change of HTC for the metal-mold contact region is not
that remarkable so far the growth rate of the solid shell is concerned for this melt feeding system and for a higher casting speed
up to a certain value of HTC the above observation is true.

6. Conclusions
The steady state phase of DCC process for an industrial-sized
rolling ingot with a combo-bag having a bottom porous foam
plate is simulated for the rst time. The following observations are
found:
[1] For an increase in casting speed a deeper sump, a thinner
shell, a higher surface temperature, and a greater level of
turbulence are attained due to the increased intensity of
thermal convection from the incoming melt.
[2] A thicker mushy region is observed at the ingot center with
the increase of the casting speed.
[3] For the studied melt feeding system two vortices are seen to
develop. One vortex forms in the bottom region and right
after the exit of the side window of the combo bag. Another
vortex forms at the ingot top free surface. With the increase
of the casting speed, the strength of these two vortices
increases.
[4] It is observed for the rst time that at the narrow symmetry
plane the melt is owing from the rolling face towards the
ingot center in an inclined manner, termed here as an angle
ow. The strength of the angle ow is increasing with the
increase of the casting speed.
[5] The vertical jet emerging from the bottom of the porous plate
of the combo bag near the central region is colliding with the
angle ow and the strength of the downward jet is reduced.
This has resulted in the attening of the solidication front in
the central area. With the increase of the casting speed the
extent of the atness of the solidication front in the central
region has progressively increased. Because of the above
colliding effects of the jet and angle ow there is only a
minor effect on the enhancement of the sump depth with
casting speed.
[6] For the rst time for a slab caster a cliff-shaped mushy region
is predicted right after the exit of the side window of the bag.
This is due to the enhanced thermal convection currents
resulting from the merging of the angle ow and the vortex
that forms right after the side window of the distribution
bag.
[7] With the increase of the casting speed the cliff-shaped
mushy region is moving downward and is taking different
shapes due to the stronger thermal convection.
[8] For the studied casting speeds, a thicker solid shell is predicted at the mid-point of the wide face compared to the
mid-point of the narrow face at the exit of the mold.

86

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87

[9] After a certain critical value, an increase in the HTC at the


metal-mold contact region has resulted in a signicant increase in the shell thickness for the higher value of the
casting speed, while at the lower casting speed there is an
insignicant effect on the shell thickness for the change in
HTC.
Acknowledgments
This work is partially supported from the National Sciences and
Engineering Research Council (NSERC) of Canada Discovery Grant #
48158 awarded to M. Hasan of McGill University, Montreal, for
which authors are grateful.
Nomenclature
A
ap, anb, b
c1, c2, cm
cp
CF
D
Dk
Da
E
f1, f2, fm
fl
fs
G
Gr
Grm
gx
h
H
K0
k
kuid
ksolid
keffective
P
Pe
Re
ReP
Ret
S
SF
Sk*
S*
Ste
T
T0
Tin
Tl
Ts
Tsurf.
ui
ui
uin
us
U, V, W
Us

Darcy coefcient
coefcients in the discretized governing equations
empirical constants for low Reynolds number model
specic heat
morphology constant
nozzle hydraulic diameter
extra dissipation term in k-equation
Darcy number
extra generation term in -equation
empirical constants used in low-Re version of ke models
liquid fraction
solid fraction
production term in turbulent kinetic energy equation
Grashof number
modied Grashof number Gr/Ste
gravitational acceleration in the x-direction
sensible heat
total heat (sensible and latent)
permeability of porous media
turbulent kinetic energy
thermal conductivity of the liquid
thermal conductivity of the solid
effective thermal conductivity of the porous media
hydrodynamic pressure
Peclet number
Reynolds number
pore Reynolds number Re/f
turbulent Reynolds number based on the turbulent
quantities
source term
source term associated with F
non-dimensional source term associated with buoyancy
term in k-equation
non-dimensional source term associated with buoyancy
term in -equation
Stefan number
temperature
uctuation of temperature
inlet temperature
liquidus temperature
solidus temperature
surface temperature of the caster
velocity component in the i-th direction; corresponding
to u
time-average velocity component in the i-th direction
inlet velocity
casting speed
non-dimensional form of the u, v and w velocities
non-dimensional form of us

x
y
z
X, Y, Z

axial direction
horizontal direction parallel to the wide face
horizontal direction parallel to the narrow face
non-dimensional form of x, y, z

Greek symbols
nodal latent heat
latent heat of solidication
diffusion coefcient associated with F
alloy density
effective viscosity equal to mt ml
laminar viscosity
turbulent viscosity or eddy diffusivity
generalized dependent variable
effective diffusivity

rate of energy dissipation


in
inlet rate of energy dissipation
g
effective convective heat transfer coefcient
sk, s
turbulence model constants
st
turbulent Prandtl number
f
porosity of lter

DH
DHf
GF
r
me
ml
mt
F
Geff

Superscripts
*
non-dimensional variables
e
time-averaged variables
0
uctuation of variables.
References
[1] A. Prasad, I.F. Bainbridge, Experimental determination of heat transfer across
the metal/mold gap in a direct-chill (DC) casting mold-part-I: effect of gap size
and mold gas type, Metall. Mater. Trans. A 44A (2013) 456e468.
[2] J. Sengupta, B.G. Thomas, M.A. Wells, The use of water cooling during the
continuous casting of steel and aluminum alloys, Metall. Mater. Trans. A 36A
(2005) 187e204.
[3] J.F.R.C. Etienne, A.R. Baserinia, H. Ng, M.A. Wells, D.C. Weckman, Heat-transfer
measurements in the primary cooling phase of the direct-chill casting process,
Metall. Mater. Trans. B 43B (2012) 1202e1213.
[4] E.K. Jensen, W. Schneider, Investigations About Starting Cracks in DC-Casting
of 6063-Type Billets, II. Modeling Results, Light Metals, The Minerals, Metals
and Materials Society, Warrendale, PA, 1990, pp. 937e943.
[5] J. Jacoby, Direct chill casting defects, aluminum cast house technology, in: 5th
Australian Asian Pacic Conference, 1995.
[6] J.F. Grandeld, V. Nguyen, et al., Ingot caster productivity improvement
through examination of mould heat ow and deformation, in: Proceedings of
the Australasian Conference and Exhibition, Aluminum Cast House Technology, Sydney, 2007, pp. 147e154.
[7] D.G. Eskin, Physical metallurgy of direct chill casting of aluminum alloys, in:
J.N. Fridlyander, D.G. Eskin (Eds.), Advances in Metallic Alloys, CRC Press, 2008.
[8] S.P. Tremblay, M. Vincent, Development and Use of a New Composite Material
for Aluminium Contact Applications, Light Metals, The Minerals, Metals and
Materials Society, Warrendale, PA, 2005, pp. 1007e1012.
[9] S.P. Tremblay, M. Lapointe, The Manufacturing, Design and Use of a New
Reusable Molten Metal Distributor for Sheet Ingot Casting, Light Metals, The
Minerals, Metals and Materials Society, Warrendale, PA, 2002, pp. 961e965.
[10] R. Kieft, J. van Oord, F. Frinking, D. Bal, H. van Schoonevelt, Detailed Modeling
of a Metal Distributor by Means of a Combined Numerical and Physical
Approach, Light Metals, The Minerals, Metals and Materials Society, Warrendale, PA, 2003, pp. 821e827.
[11] M. Fortier, A. Larouche, X.G. Chen, Y. Caron, The Effect of Process Parameters
on the Metal Distribution for DC Sheet Ingot Casting, Light Metals, The Minerals, Metals and Materials Society, Warrendale, PA, 2005, pp. 1019e1024.
[12] A. Arsenault, D. Larouche, S.P. Tremblay, J.P. Dube, DC Cast Thermal and Fluid
Flow Simulation Using a Semi-permeable Model of TF Combo Bag, Light
Metals, The Minerals, Metals and Materials Society, Warrendale, PA, 2008, pp.
781e785.
[13] F. Ilinca, J.F. Hetu, A. Arsenault, D. Larouche, S.P. Tremblay, 3-D Modeling of
the Flow and Heat Transfer During DC Casting with a Combo Bag, Light Metals,
The Minerals, Metals and Materials Society, Warrendale, PA, 2009, pp.
799e804.
[14] D.R. Poirier, G.H. Geiger, Transport Phenomena in Materials Processing, The
Minerals, Metals and Materials Society, Warrendale, PA, 1994.
[15] J. Zuidema Jr., L. Katgerman, I.J. Opstelten, J.M. Rabenberg, Secondary Cooling
in DC Casting: Modeling and Experimental Results, Light Metals, The Minerals,
Metals and Materials Society, Warrendale, PA, 2001, pp. 873e878.

L. Begum, M. Hasan / International Journal of Thermal Sciences 86 (2014) 68e87


[16] B.E. Launder, B.I. Sharma, Application of the energy dissipation model of
turbulence to the calculation of the ow near a spinning disk, Lett. Heat Mass
Transfer 1 (1974) 131e138.
[17] L. Begum, 3-D Transport Phenomena in Vertical Direct Chill Casting Processes
(Ph.D. thesis), McGill University, Montreal, Quebec, Canada, 2013.
[18] V.R. Voller, C. Prakash, A xed grid numerical modeling methodology for
convectionediffusion mushy region phase-change problems, Int. J. Heat Mass
Transfer 30 (1987) 1709e1719.
[19] S.H. Seyedein, M. Hasan, A three-dimensional simulation of coupled turbulent
ow and macroscopic solidication heat transfer for continuous slab casters,
Int. J. Heat Mass Transfer 41 (1997) 4405e4423.
[20] S. Minakawa, I.V. Samarasekera, F. Weinberg, Centerline porosity in plate
castings, Metall. Trans. 16B (1987) 245e255.
[21] C.J. Vreeman, F.P. Incropera, The effect of free-oating dendrites and convection on macrosegregation in direct chill cast aluminum alloys part II:
predictions for AleCu and AleMg alloys, Int. J. Heat Mass Transfer 43 (2000)
687e704.
[22] C. Beckermann, S. Ramadhyani, R. Viskanta, Natural convection ow and heat
transfer between a uid layer and a porous layer inside a rectangular enclosure, J. Heat Transfer Trans. ASME 109 (2) (1987) 363e370.
[23] C. Beckermann, R. Viskanta, S. Ramadhyani, A numerical study of non-Darcian
natural convection in a vertical enclosure lled with a porous medium,
Numer. Heat Transfer 10 (6) (1986) 557e570.
[24] M.H.J. Pedras, M.J.S. de Lemos, On the denition of turbulent kinetic energy for
ow in porous media, Int. Commun. Heat Mass Transfer 27 (2) (2000)
211e220.
[25] M.H.J. Pedras, M.J.S. de Lemos, Macroscopic turbulence modeling for incompressible ow through undeformable porous media, Int. J. Heat Mass Transfer
44 (2001a) 1081e1093.
[26] M.H.J. Pedras, M.J.S. de Lemos, On the mathematical description and simulation of turbulent ow in a porous medium formed by an array of elliptic rods,
J. Fluids Eng. 123 (2001b) 941e947.
[27] M.H.J. Pedras, M.J.S. de Lemos, Computation of turbulent ow in porous media
using a low-Reynolds ke model and an innite array of transversally displaced elliptic rods, Numer. Heat Transfer Part A 43 (2003) 585e602.
[28] K. Vafai, C.L. Tien, Boundary and inertia effects on ow and heat transfer in
porous media, Int. J. Heat Mass Transfer 24 (1981) 195e203.

87

[29] J.T. Hong, C.L. Tien, M. Kaviany, Non-Darcian effects on vertical plate natural
convection in porous media with high porosities, Int. J. Heat Mass Transfer 28
(11) (1985) 2140e2157.
[30] J.L. Large, The fundamental theory of ow through permeable media from
Darcy to turbulence, in: D.B. Ingham, I. Pop (Eds.), Transport Phenomena in
Porous Media, Elsevier Science, 1998, ISBN 0-08-042843-6, p. 446.
[31] Y.T. Yang, M.L. Hwang, Numerical simulation of turbulent uid ow and heat
transfer characteristics in a rectangular porous channel with periodically
spaced heated blocks, Numer. Heat Transfer Part A 54 (2008) 819e836.
[32] K. Vafai, R.L. Alkire, C.L. Tien, An experimental investigation of heat transfer in
variable porosity media, J. Heat Transfer Trans. ASME 107 (3) (1985) 642e647.
[33] S.V. Patankar, Numerical Heat Transfer and Fluid Flow, rst ed., Hemisphere
Publ. Corp., Taylor and Francis Group, New York, 1980, pp. 88e90, pp.
126e131.
[34] S.V. Patankar, Numerical prediction of three-dimensional ows, in:
B.E. Launder (Ed.), Studies in Convection: Theory, Measurement and Applications, Vol. 1, Academic Press, New York, 1975, pp. 125e127.
[35] H.K. Versteeg, W. Malalasekera, An Introduction to Computational Fluid Dynamics, the Finite Volume Method, rst ed., Longman Scientic Technical,
Longman Group Ltd., Essex, England, 1995, pp. 142e146.
[36] W.K. Jones Jr., D. Xu, J. Evans, W.E. Williams, D.P. Cook, Effects of Combo-bag
Geometry on the Thermal History and Sump Prole of a 3104 DC Cast Ingot,
Light Metals, The Minerals, Metals and Materials Society, Warrendale, PA,
1999, pp. 841e845.
[37] M. Can, A.B. Etemoglu, Porosity measurement of stainless steel lters produced by electrical discharge technique, Filtr. Sep. 41 (2004) 37e40.
[38] T.C. Dickenson, Filters and Filtration Handbook, Elsevier Advanced Technology, Oxford, U.K., 1992, pp. 123e126.
[39] L. Zhang, D.G. Eskin, A. Miroux, T. Subroto, L. Katgeran, Inuence of melt feeding
scheme and casting parameters during direct-chill casting on microstructure of
an AA7050 billets, Metall. Mater. Trans. B 43B (2012) 1565e1573.
[40] J.F. Grandeld, D.G. Eskin, I. Bainbridge, Direct-chill Casting of Light Alloys:
Science and Technology, Wiley, Hoboken, NJ, 2013, p. 244.
[41] A. Prasad, I.F. Bainbridge, Experimental determination of heat transfer with
the metal/mold gap in a direct-chill (DC) casting mold: part-II. Effect of casting
metal, mold material, and other casting parameters, Metall. Mater. Trans. A
44A (2013) 3099e3113.

Anda mungkin juga menyukai