Вы находитесь на странице: 1из 13

Journal of Alloys and Compounds 319 (2001) 174186

www.elsevier.com / locate / jallcom

Determination of transient interfacial heat transfer coefficients in chill


mold castings
C.A. Santos, J.M.V. Quaresma, A. Garcia*
Department of Materials Engineering, State University of Campinas -UNICAMP, PO Box 6122, 13083 -970 Campinas, SP, Brazil
Received 29 February 2000; accepted 22 December 2000

Abstract
The present work focuses on the determination of transient moldenvironment and metalmold heat transfer coefficients during
solidification. The method uses the expedient of comparing theoretical and experimental thermal profiles and can be applied both to pure
metals and metallic alloys. A solidification model based on the finite difference technique has been used to provide the theoretical results.
The experiments were carried out by positioning the thermocouples in both metal and mold. The comparison between experimental and
theoretical results is made by an automatic search of the best fitting among theoretical and experimental cooling curves simultaneously in
metal and in mold. This has permitted the evaluation of the variation of heat transfer coefficients along the solidification process in
unsteady state unidirectional heat flow of AlCu and SnPb alloys, as well as the analysis of the effects of the material and the thickness
of the mold and melt superheat. 2001 Elsevier Science B.V. All rights reserved.
Keywords: Interfacial heat transfer coefficients; Mathematical modeling; Chill mold casting; AlCu; SnPb

1. Introduction
The structural integrity of shaped castings is closely
related to their temperaturetime evolution during solidification. A number of analytical and numerical models were
developed in the last 2 decades to treat heat transfer during
solidification, and the consequent simulation of freezing
patterns in castings has provided many improvements in
foundry processes. The use of casting solidification simulation could do much to increase knowledge of the process,
however, some uncertainties must be eliminated before
such simulations can be widely accepted as realistic
descriptions of the process. The heat transfer at the metal
mold interface is one of these uncertainties, and reliable
experimental values of heat transfer coefficients are required for various metalmold combinations and superheats, as existing data is sparse. The way the heat flows
across the metal and mold surfaces directly affects the
evolution of solidification, and plays a notable role in
determining the freezing conditions within the metal,
mainly in foundry systems of high thermal diffusivity like
chill castings. Gravity or pressure die casting, continuous
casting, and squeeze casting are some of the processes
*Corresponding author.
E-mail address: amaurig@fem.unicamp.br (A. Garcia).

where product quality is more directly affected by the


interfacial heat transfer conditions. Once information in
this area is accurate, foundrymen can effectively optimize
the design of their chilling systems to produce sound
castings.
When metal and mold surfaces are brought into contact
an imperfect junction is formed. While uniform temperatures gradients can exist in both metal and mold, the
junction between the two surfaces creates a temperature
drop, which is dependent upon the thermophysical properties of the contacting materials, the casting and mold
geometry, the roughness of mold contacting surface, the
presence of gaseous and non-gaseous interstitial media, the
melt superheat, contact pressure and initial temperature of
the mold.
Fig. 1a shows a schematic representation of the two
contacting surfaces. Because the two surfaces in contact
are not perfectly flat, when the interfacial contact pressure
is reasonably high, most of the energy passes through a
limited number of actual contact spots [1,2]. The heat flow
across the castingmold interface can be characterized by
a macroscopic average metalmold interfacial heat transfer
coefficient (h i ), given by
q
h i 5 ]]]]
A(T IC 2 T IM )

0925-8388 / 01 / $ see front matter 2001 Elsevier Science B.V. All rights reserved.
PII: S0925-8388( 01 )00904-5

(1)

C. A. Santos et al. / Journal of Alloys and Compounds 319 (2001) 174 186

Fig. 1. Heat flow across moldcasting interface.

where q (W) is the average heat flux across the interface


and T IC and T IM are, respectively casting and mold surface
temperatures (K).
The quantification of heat flux in terms of a heat transfer
coefficient, as indicated in Fig. 1b for the idealized
temperature profile, requires that the heat capacity is zero
so that the thermal diffusivity is infinite, and consequently
heat fluxes entering and leaving the interface are equal.
The heat transfer coefficient shows a high value in the
initial stage of solidification, the result of the good surface
conformity between the liquid core and the solidified shell.
As solidification progresses the mold expands due to the
absorption of heat and the solid metal shrinks during
cooling and as a result a gap develops because pressure
becomes insufficient to maintain a conforming contact at
the interface. Once the air gaps forms, the heat transfer
across the interface decreases rapidly and a relatively
constant value of h i is attained. The mode of heat transfer
across the metalmold interface has been suggested to be
due to both conduction through isolated metalmold
contacts and through gases present in the gap and radiation
between the surfaces. During the subsequent stage of
solidification a slight drop in the interfacial heat transfer
coefficient with time can be observed. It is postulated that
this is caused by the growth of oxide films on chill and
mold surfaces [3], and by a reduction in the thermal

175

conductivity of the interfacial gas with declining temperature.


By using measured temperatures in both casting and
mold, together with numerical [410] or analytical [11,12]
solutions of the solidification problem, many research
workers have attempted to quantify metalmold interfacial
heat transfer in terms either of a heat transfer coefficient or
heat flux. In most cases, the numerical techniques generally known as the method of solving the inverse heat
conduction problem were used to quantify the time dependent heat transfer coefficient at the interface. Growth data
obtained from the dendritic microstructure together with a
numerical solution have also been used to determine
metalmold transient heat transfer coefficients [13].
In the present study, the heat flow between the casting
and the chill is characterized by the interfacial heat transfer
coefficient, h i . The variation of h i during solidification of
AlCu and SnPb alloys, as well as pure aluminum and
pure tin, against a vertical mold wall is investigated
experimentally. The effects of mold material (low carbon
steel and copper) and its thickness and melt superheat are
also investigated. Experimental temperatures in the mold
and the metal during solidification are compared with
simulations furnished by a numerical model, and an
automatic search selects the best theoreticalexperimental
fitting from a range of values of h i . In any case examined,
expressions are derived representing variation of metal
mold and moldenvironment heat transfer coefficients with
time.

2. Determination of interfacial heat transfer


coefficients

2.1. Metalmold heat transfer coefficient h i


The heat flow across the castingmold interface can be
characterized by Eq. (1) and h i can be determined provided that all the other terms of the equation, namely q,
T IC and T IM , are known. However, these temperatures are
difficult to measure because the accurate location of
thermocouples of finite mass at the interface is not an easy
task, and they can distort the temperature field at the
interface. To overcome this experimental impediment, the
methods of calculation of h i existing in the literature are
based on a knowledge of other conditions, such as
temperature histories at interior points of the casting or
mold, together with mathematical models of heat flow
during solidification. Among these methods, those based
on the solution of the inverse heat conduction problem
have been widely used in the quantification of the transient
interfacial heat transfer. Since solidification of a casting
involves both a change of phase and temperature variable
thermal properties, the inverse heat conduction becomes
nonlinear. The nonlinear estimation technique was used by
Beck for the numerical solution of this class of problem

C. A. Santos et al. / Journal of Alloys and Compounds 319 (2001) 174 186

176

[14]. It has advantage over the other numerical procedures,


in that Beck studied the problem from the standpoint of
effective treatment of experimental data, taking into account inaccuracies concerning the locations of thermocouples, statistical errors in temperature measurement and
uncertainty in material properties.
In the present work, a similar procedure determines the
value of h i which minimizes an objective function defined
by the equation

O (T
n

F(h) 5

2
est 2 T exp )

(2)

i 51

where T est and T exp are, respectively, the estimated and the
experimentally measured temperatures at various thermocouples locations and times, and n is the iteration stage.
A suitable initial value of h i is assumed and with this
value, the temperature of each reference location in casting
and mold at the end of each time interval Dt is simulated
by using an explicit finite difference technique. The
correction in h i at each iteration step is made by a value
Dh i , and new temperatures are estimated [T est (h i 1 Dh i ) or
T est (h i 2 Dh i )]. With these values, sensitivity coefficients
(f ) are calculated for each iteration, given by

T est (h i 1 Dh i ) 2 T est (h i )
f 5 ]]]]]]]
Dh i

(3)

The sequence of the solution involves the calculation of


the sensitivity coefficients for measured temperatures. The
assumed value of h i is corrected using the relation
h i (new) 5 h i (old)6Dh i

(4)

The above-indicated procedure is repeated for a new


value of h i , and is continued until
Dh
]i , 0.01
hi

(5)

The calculation of h i as a function of time is continued


until the end of the desired period. The flow chart shown in
Fig. 2 gives an overview of the solution procedure.

2.2. Heat flow model


With adequate insulation of the chill and casting
chamber, heat flow through the casting can be reasonably
approximated as a one-dimensional heat transfer problem,
which can be analyzed by
T
2T
r c ] 5 k ]2 1 q~
t
x

(6)

where r, c, k are, respectively, density (kg / m 3 ), specific


heat (J / kg K) and thermal conductivity (W/ m K), T is
temperature, t is time (s) and x is distance along the x axis
(m). The term q~ on the right hand side of Eq. (6) is a heat
source term which is incorporated to account for the latent
heat of solidification, and is given by
fS
q~ 5 r L ]
t

(7)

where L is the latent heat of fusion (J / kg) and fS is the


solid fraction. When treating the chill heat flow, the
governing equation is similar to Eq. (6) expect that the q~
term is not included.
Eq. (7) can be related to temperature as follows
f
f T
]S 5 ]S ? ]
t
T t

(8)

Substitution of Eqs. (7) and (8) into Eq. (6) gives

S D

fS T
2T
r c 2 L ? ] ] 5 k ]2
T t
x

(9)

fS
The term (L ? ] ) in Eq. (9) can be considered as a pseudo
T
specific heat and an apparent specific heat (c9) can be
defined, and this equation can be written as
Fig. 2. Flow chart for the determination of metalmold heat transfer
coefficients.

S D

T
2T
r c9] 5 k ]2
t
x

(10)

C. A. Santos et al. / Journal of Alloys and Compounds 319 (2001) 174 186

The introduction of finite difference terms to Eq. (10)


gives

D S

n
T ni 11 2 T ni
T i21
2 2T ni 1 T ni11
r c9 ]]] 5 k ]]]]]]
Dt
Dx 2

(11)

where the subscripts indicate the node address on the


spatial network and the superscripts represent time. Multiplying Eq. (11) by (Dx Dy Dz) yields

n
n
T in11 2 T in
T i11
1 T i21
2 2T in
A T Dxr c9 ]]] 5 A T k ]]]]]]
Dt
Dx

where A T 5 Dy Dz (m ), and y and z are the distances,


respectively, along y axis and z axis. By applying an
analogy between electrical and thermal circuits, the energy
accumulated in a volume element i, is given by
(13)

where V is the finite volume element (m 3 ), and CTi is the


thermal capacitance (J / kg).
The thermal resistance at the heat flux line x can be
calculated for each element, and given by
Dx i11
R i11 5 ]]]
2k i11 A T

(14)

Dx i21
R i21 5 ]]]
2k i21 A T

(15)

Dx i
R i 5 ]]
2k i A T

(16)

CTi

DS

DS

(17)

or

S D

Dt
Dt
T n11
5 ] ? T ni11 1 1 2 ]]
i
tQi
tQDi

The heat transfer coefficients at the moldair interface


can be calculated as a function of measured mold wall
temperatures (T EM ) and free-stream air temperature (T 0 )
[17]. The thermal resistance at this interface is given by
1
R M /A 5 ]]]]
(h R 1 h C )A T

h R 5 se (T EM 1 T 0 ) ? (T 2EM 1 T 20 )

(22)

S D

Dt
n
? T in 1 ] ? T i21
tDi

(23)

where s is the StefanBoltzmann constant (5.672310


W/ m 2 K 4 ) and e is the mold emissivity.
The convection heat transfer coefficient is given by
k gas
Nu
h C 5 ]]
x

28

(24)

where h C are represented in terms of Nusselt number (Nu ),


and for free convection can be calculated as a function of
Grashof (GR ) and Prandtl (PR ) numbers, as follows [15]
Nu 5 C (GR PR )n

(25)

where C and n are constants, and x is a characteristic


length of the solid surface (m), in our particular case the
chill vertical length. GR and PR , are given, respectively, by
ggx 3 (T EM 2 T 0 ) 2
GR 5 ]]]]]
? rs
h2

(26)

h
PR 5 ] ? c
k

(27)

Introducing Eqs. (1316) into Eq. (12) yields


n
T n11
2 T ni
T i11
2 T ni
T ni21 2 T ni
i
]]]
5 ]]] 1 ]]]
Dt
R i11 1 R i
R i21 1 R i

2.3. Moldenvironment heat transfer coefficient h a

where h R and h C are, respectively, the radiation and


convection heat transfer coefficients, calculated as follows
(12)

CTi 5 A T Dx i ri c 9i 5Vri c 9i

177

where g is the gravitational acceleration (m / s 2 ), g is the


volume coefficient of expansion [for ideal gases g 5 1 /T 0
(K 21 )], h is fluid viscosity, r is fluid density, k is fluid
thermal conductivity and c is the fluid specific heat.

3. Experimental

(18)
where

tQi 5 CTi (R i11 1 R i )

(19)

tDi 5 CTi (R i21 1 R i )

(20)

tQitDi
tQDi 5 ]]]
tQi 1 tDi

(21)

Eq. (18) represents the solution of the explicit form of


the finite difference method, and will be stable for Dt #
tQDi . A three dimensional version of this solution has been
recently applied for cases of complex shaped bodies
[15,16].

Experiments were performed with Sn, Pb, Al and SnPb


and AlCu alloys, including short and long freezing range
alloys, as well as eutectic compositions. The casting and
chill materials selected for experimentation, and the employed thermophysical properties are summarized in Table
1.
The casting assembly used in solidification experiments
is shown in Fig. 3. The main design criteria were to ensure
a dominant unidirectional heat flow during solidification.
This objective was achieved by adequate insulation of the
chill and casting chamber.
Copper and a low carbon steel chills were used, with the
heat-extracting surfaces being polished. In order to investigate the influence of chill thickness on heat flow, four

C. A. Santos et al. / Journal of Alloys and Compounds 319 (2001) 174 186

178

Table 1
Casting and chill materials used for experimentation and the corresponding thermophysical properties [1822] a

k S (W/ m K)
k L (W/ m K)
c S (J / kg K)
c L (J / kg K)
rS (kg / m 3 )
rL (kg / m 3 )
aS (m 2 / s)
aL (m 2 / s)
L (J / kg)
T F (8C)
T E (8C)
T L (8C)
e
K0
a

Al

Al
4.5%Cu

Al
15%Cu

Al
33%Cu

Steel
SAE
1010

Pb

Sn
39%Pb

Sn
20%Pb

Sn
10%Pb

Sn
5%Pb

Sn

Copper

222
92
1123
1086
2550
2380
7.75(310 25 )
3.36(310 25 )
385 000
660

193
85
1092
1059
2650
2480
6.67
3.24
381 900
660
548
645

179
80
1080
999
2910
2760
5.67
2.90
274 270
660
548
618

155
71
1070
895
3410
3240
4.25
2.45
350 000
660
548

46

34.7
29.7
129.8
138.2
11 340
10 678
2.37
2.04
26 205
327

54.7
31.7
186.2
212.9
8840
8400
3.35
1.79
47 560
232
183

59
32
200
231
8250
7860
3.58
1.76
52 580
232
183
202

63
33
209
243
7840
7480
3.84
1.81
56 140
232
183
216

64
33
221
259
7720
7380
3.91
1.82
57 120
232
183
220

67
33
221
259
7300
7000
4.15
1.82
60 710
232

372

0.17

0.17

0.0656

0.0656

0.0656

527
7860

0.8

419
8960

0.023

a, Thermal diffusivity; T E , eutetic temperature, T L , liquidus temperature; Ko , partition coefficient.

different thicknesses of chills were used (X57, 17, 28, 39


and 60 mm).
Two chromelalumel thermocouples were introduced in
the chill; one near the chillcasting interface and the other
at the outer surface, and a third one was placed in the

casting and located 20 mm from the interface, as indicated


in Fig. 3. All of the thermocouples were connected by
coaxial cables to a data logger interfaced with a computer,
and the temperature data were acquired automatically. The
temperature files were used in a finite-difference heat flow
program to estimate the transient heat transfer coefficients.
A schematic representation of the experimental setup
connected to the data acquisition and analysis system is
shown in Fig. 4.
Each alloy was melted in an electric resistance-type
furnace until the molten metal reached a predetermined
temperature. It was then stirred until the temperature was
brought to a specified value and poured into the casting
chamber. The aluminum alloys were degassed with hexachloroethane tablets before pouring.
The effect of liquid metal superheat on heat transfer
coefficient was also investigated, by using a Sn 10%Pb
alloy, a 60-mm thick carbon steel chill and different
degrees of superheat: 20, 40, 70 and 1008C above liquidus
temperature. The thermocouples were calibrated at the
melting points of aluminum (for AlCu alloys) and tin (for
SnPb alloys), exhibiting fluctuations of about 1.0 and
0.48C, respectively. The experimental profiles plotted are
the averages of three thermocouple readings at each
location in chill and casting. Results from repeated experiments have shown differences not greater than 48C, as can
be seen in Fig. 5 for the case of pure Sn.

4. Results and discussion

4.1. Moldenvironment heat transfer coefficients h a

Fig. 3. Experimental set-up (mm).

Figs. 6 and 7 show typical examples of temperature data


collected in chill during the course of solidification experiments with a Sn 10 wt.% Pb alloy. In both cases of carbon

C. A. Santos et al. / Journal of Alloys and Compounds 319 (2001) 174 186

179

Fig. 4. Schematic representation of the experimental setup connected to the data acquisition and analysis system.

steel and copper chills, a similar trend can be observed.


For the smaller chill thicknesses, the external surface
temperature rises rapidly from the beginning of solidification until a peak value, and declines thereafter. On the
other hand, a progressive rise is observed when the thicker
chills are used. It is obvious that the moldenvironment
heat transfer coefficients are expected to follow the same
trend.
The resulting transient values of h a were calculated by

using the above mentioned measured surface temperatures,


Eqs. (2227) and the thermophysical properties of Table 1.
For the 6-mm thick chills (carbon steel and copper) a
progressive decrease in h a were observed for t . 800 s, but
the differences in h a were so small during the course of
solidification that it was assumed constant and equal to the
peak value for t . 800 s.
The effect of thickness of mold was experimentally
investigated, and the typical results are shown in Fig. 8. No
significant differences in h a can be observed, when the
steel mold is replaced by a massive copper chill. Similar
results were obtained during experimental investigation of
AlCu alloys, so that for purposes of numerical simulations performed in next section for the determination of
metalmold heat transfer coefficients, the same experimental equation has been adopted for each metallic system
SnPb h a 5 5.7 t 0.15

(28)

AlCu h a 5 5.1 t 0.27

(29)
2

where t (s) and h a (W/ m K).

4.2. Metalmold heat transfer coefficients h i

Fig. 5. Repeated experiments for pure Sn.

4.2.1. Effect of alloy composition


Solidification simulation of each test casting was performed by adopting two different approaches for the
liberation of the latent heat of fusion. For eutectic alloys
and pure metals, the latent heat (L) was transformed into
the equivalent number of degrees by considering a temperature accumulation factor ( l) related to L by the
specific heat ( l 5 L /c). For short or long freezing range
alloys, the latent heat evolution was taken into account by
using Scheils equation until the remaining liquid reached

180

C. A. Santos et al. / Journal of Alloys and Compounds 319 (2001) 174 186

Fig. 6. Typical experimental temperature responses in steel molds at two locations: external wall temperature and at 3 mm from metalchill interface: (A)
6 mm mold thick and (B) 60 mm mold thick.

the eutectic composition. Temperature was experimentally


measured in two locations: in the chill at 3 mm from the
metalmold interface and in the casting at 20 mm from
this interface. In Figs. 9 and 10 typical experimental

thermal responses are compared to those numerically


simulated by using the transient h i profile which provides
the best curve fitting.
Figs. 11 and 12 show the metalmold heat transfer

Fig. 7. Typical experimental temperature responses in cooper molds at two locations: external wall temperature and at 3 mm from metalchill interface:
(A) 6 mm mold thick and (B) 60 mm mold thick.

C. A. Santos et al. / Journal of Alloys and Compounds 319 (2001) 174 186

181

Fig. 8. Calculated moldenvironment heat transfer coefficients h a : (A) steel molds with different thicknesses, (B) copper molds with different
thicknesses.

coefficients profiles as a function of time, respectively for


the cases of SnPb and AlCu alloys solidifying against a
60 mm thick carbon steel chill. The observed differences in
the h i profiles between the pure metals and the other alloys
examined, can be explained by the total shrinkage accompanying solidification, the extent of the solidification range
and the wetting of the mold by the melt. For both metallic
systems, the h i profile increases with increasing mushy

zone length. For longer mushy zones, the interdendritic


liquid can feed better the solidification contraction causing
a continued presence of liquid at the interface, leading to
higher values of h i . This can be taken as a general trend,
but care should be exercised when applying this conclusion
to the beginning of solidification.
As can be seen in Fig. 12, the Al 15 wt.% Cu alloy
exhibits initial h i values higher than those corresponding to

182

C. A. Santos et al. / Journal of Alloys and Compounds 319 (2001) 174 186

Fig. 9. Typical experimental temperature responses at two locations in casting and chill: in casting at 20 mm from the metalmold interface and in chill at
3 mm from this interface: (A) Sn 10 wt.% Pb, (B) pure Sn both solidified in a 60 mm thick steel chill and a superheat DT 5 0.1 T L (10% of liquidus or
melting temperatures).

the Al 4.5 wt.% Cu alloy, which has a longer mushy zone.


At the initial stage of solidification the wetting of the mold
by the melt seems to be the dominant factor controlling

heat transfer coefficient. Anyway, a more complex experimental set-up and a numerical technique dealing with
convection heat transfer would be necessary for an accur-

Fig. 10. Typical experimental temperature responses at two locations in casting and chill: in casting at 20 mm from the metalmold interface and in chill at
3 mm from this interface: (A) Al 4.5 wt.% Cu, (B) pure Al both solidified in a 60 mm thick steel chill and a superheat DT 5 0.1 T L (A) and DT 5 0.2 T L (B).

C. A. Santos et al. / Journal of Alloys and Compounds 319 (2001) 174 186

183

Fig. 11. Evolution the metalmold interfacial heat transfer coefficients as a function of alloy composition: SnPb system, 60 mm thick steel chill and a
superheat DT 5 0.1 T L .

Fig. 12. Evolution the metalmold interfacial heat transfer coefficients as a function of alloy composition: AlCu system, 60 mm thick steel chill and a
superheat DT 5 0.1 T L (10% of liquidus or melting temperatures).

184

C. A. Santos et al. / Journal of Alloys and Compounds 319 (2001) 174 186

Fig. 13. Evolution of the metalmold heat transfer coefficients as a function of chill thickness: Sn 10 wt.% Pb alloy, steel chill and a superheat DT 5 0.1
TL.

ate characterization of initial heat transfer coefficients, as


fluid flow in the cast alloys, with its associated heat
transfer, would be at its strongest.

4.2.2. Effects of chill material and chill thickness


The effects of chill material and chill thickness on heat
transfer coefficient are shown in Figs. 13 and 14, for the
cases of a Sn 10 wt.% Pb alloy solidifying, respectively

against a carbon steel and copper chills at a superheat of


DT 5 0.1 T L . As can be seen by comparing Figs. 12 and
13, the heat transfer coefficient profiles increase with
increasing thermal diffusivity of the chill material. These
results are in agreement with other studies in the literature
[4,10]. The h i profiles increase with decreasing chill
thickness. The chill temperature rises more rapidly from
the beginning of solidification with decreasing chill thick-

Fig. 14. Evolution of the metalmold heat transfer coefficients as a function of chill thickness: Sn 10 wt.% Pb alloy, cooper chill and a superheat DT 5 0.1
TL.

C. A. Santos et al. / Journal of Alloys and Compounds 319 (2001) 174 186

185

Fig. 15. Evolution of the metalmold heat transfer coefficients as a function of superheat: Sn 10 wt.% Pb alloy, 60 mm thick steel chill.

ness. As a consequence, mold expansion favors the thermal


contact between metal and chill surface and as the
solidified shell is not so thick as for thicker chills, this
translates to lower contraction away from the chill. Both
factors will contribute to an increase in h i values.

4.2.3. Effect of superheat


The heat transfer coefficient increases with increasing
values of superheat, as can be seen in Figs. 15 and 16,
respectively, for a Sn 10 wt.% Pb alloy and pure Al
solidifying against a 60 mm thick carbon steel chill. The

Fig. 16. Evolution of the metalmold heat transfer coefficients as a function of superheat: Al, 60 mm thick steel chill.

C. A. Santos et al. / Journal of Alloys and Compounds 319 (2001) 174 186

186

fluidity of molten alloys increase with increasing superheat, favoring the wetting of the chill by the melt [23,24].
Some results reported in the literature indicate that the
surface of solidified shell becomes smoother as the superheat increases for the same chill microgeometry, thus
increasing the interfacial contact [3]. The influence of
superheat is not so significant for the Sn 10 wt.% Pb alloy.
In fact some differences can be observed only for values
higher than 408C (Fig. 15). This is not the case for Al,
where fluidity plays a more significant role. The initial
value of h i rises from about 1500 to 6000 W/ m 2 K if the
superheat is increased from 10% of the melting point (T f )
to 20% of T f (Fig. 16).

The h i profiles increased with increasing thermal diffusivity of the chill material and with decreasing
thickness of the chill.

Acknowledgements
The authors would like to acknowledge financial support
provided by FAPESP (The Scientific Research Foundation
Paulo, Brazil) and CNPq (The Brazilian
of the State of Sao
Research Council).

References
5. Conclusions
Experiments were conducted to analyze the evolution of
moldenvironment (h a ) and metalmold (h i ) heat transfer
coefficients during solidification of SnPb and AlCu
alloys in vertical steel and copper chills. The following
conclusions can be drawn
The transient interfacial heat transfer coefficients (h i
and h a ) have been successfully characterized by using
an approach based on measured temperatures along
casting and chill, and analytical calculations (h a ) and
numerical simulations provided by a heat flow model
(h i ).
The moldenvironment heat transfer coefficient have
been expressed as a power function of time, given by
the general form
h a 5 Cm (t)0.15

[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]

[10]
[11]
[12]

[13]
2

where h a (W/ m K), t (s) and Cm is a constant which


depends on thickness of chill. No significant differences
were found when the steel chill was replaced by a
copper chill, considering the physical configuration and
processing parameters adopted in the study.
The metalmold heat transfer coefficients have also
been expressed as a power function of time, given by
the general form

[14]
[15]
[16]
[17]
[18]
[19]

h i 5 Ci (t)2n

[20]

where h i (W/ m 2 K), t (s) and Ci and n are constants


which depend on alloy composition, chill material and
superheat.

[21]
[22]
[23]
[24]

L.S. Fletcher, J. Heat Transfer 110 (1988) 1059.


K. Ho, R.D. Pehlke, Afs Trans 92 (1984) 587.
K. Ho, R.D. Pehlke, Metall. Trans. 16B (1985) 585.
C.A. Muojekwu, I.V. Samarasekera, J.K. Brimacombe, Metall.
Trans. 26B (1995) 361.
M. Krishnan, D.G.R. Sharma, Int. Comm. Heat Mass Transfer 23
(1996) 203.
A.V. Reddy, C. Beckermann, Exp. Heat Transfer 6 (1993) 111.
M.A. Taha, N.A. El-Mahallawy, A.W.M. Assar, R.M. Hammouda, J.
Mater. Sci. 27 (1992) 3467.
T.S. Prasanna, K. Narayan Prabhu, Metall. Trans. 22B (1991) 717.
J.F. Evans, D.H. Kirkwood, J. Beech, in: Modeling of Casting,
Welding and Advanced Solidification Processes, Vol. V, The Minerals, Metals and Materials Society, 1991, p. 531.
C.H. Huang, M.N. Ozisik, B. Sawaf, Int. J. Heat Mass Transfer 35
(1992) 1779.
A. Garcia, T.W. Clyne, M. Prates, Metall. Trans. 10B (1979) 773.
A. Garcia, T.W. Clyne, in: Proceedings International Conference
Solidification Technology in the Foundry and Casthouse, The
Metals Society, 1980, p. 33.
R. Caram, A. Garcia, in: Imeche Conf. Trans, Vol. 2, Mechanical
Engineers Publications, London, 1995, p. 555.
J.V. Beck, Int. J. Heat Mass Transfer 13 (1970) 703.
J.A. Spim Jr., A. Garcia, Mater. Sci. Eng. A 277 (2000) 198.
J.A. Spim Jr., A. Garcia, Numerical Heat Transfer B: Fundamentals
38 (2000) 75.
D.R. Poirier, E.J. Poirier, Heat Transfer Fundamentals For Metals
Casting, The Minerals, Metals and Materials Society, 1994.
L.F. Mondolfo, Mater. Sci. Technol. 5 (1976) 118.
R.D. Pehlke et al., Summary of Thermal Properties For Casting
Alloys and Mold Materials, University of Michigan, 1982.
Y.S. Toloukian et al., in: Thermophysical Properties of Matter, Vol.
1, IFI / Plenum, New York, 1970.
A. Bejan, Heat Transfer, Wiley, New York, 1993.
D. Bouchard, J.S. Kirkaldy, Metall. Mat. Trans. 28B (1997) 651.
M. Prates, H. Biloni, Metall. Trans. 3A (1972) 1501.
M.C. Flemings, Solidification Processing, McGraw Hill, 1974.

Вам также может понравиться