You are on page 1of 9

Experimental Thermal and Fluid Science 34 (2010) 10681076

Contents lists available at ScienceDirect

Experimental Thermal and Fluid Science


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

Application of an optimization method and experiment in inverse determination


of interfacial heat transfer coefcients in the blade casting process
Weihong Zhang *, Gongnan Xie, Dan Zhang
Engineering Simulation and Aerospace Computing (ESAC), The Key Laboratory of Contemporary Design and Integrated Manufacturing Technology,
Northwestern Polytechnical University, P.O. Box 552, 710072 Xian, Shaanxi, China

a r t i c l e

i n f o

Article history:
Received 21 December 2009
Received in revised form 18 March 2010
Accepted 20 March 2010

Keywords:
Interfacial heat transfer coefcient
Numerical prediction
Experimental measurement
Optimization method

a b s t r a c t
In order to effectively improve the numerical prediction accuracy in a blade investment casting process, a
new method is proposed to determine the interfacial heat transfer coefcient (IHTC) in a complicated
blade casting by combining the numerical prediction, optimization and limited experimental data. An
investment experiment of the blade is conducted to acquire the surface temperature of the casting and
the shell mould. Regarding the complicated mechanism of the interfacial heat transfer in the progressive
solidication, a new continuous model with three-step evolution is established for the castingmould
IHTC, and a power function is proposed to correlate the mouldenvironment IHTC with solidication
time as well. A globally convergent method is employed to search the optimal coefcients involved in
the IHTCs correlations. Results show that the predicted temperature based on proposed models agrees
well with the experimental data with the maximum deviation being less than 5.5%, and a signicant variation of the castingmould IHTC is observed. It is concluded that the prediction accuracy and efciency
associated with the optimization method can be greatly improved with the present IHTC models.
2010 Elsevier Inc. All rights reserved.

1. Introduction
It is known that the shape of the casting depends on the cavity
geometry of the metal die signicantly in the investment casting
process. An exact die prole, which generally takes into account
the various shrinkages involved in the casting process, is therefore,
important to improve the quality of net-shaped products. In this
sense, an accurate numerical simulation of the entire casting process is very helpful to realize optimal designs of the die-cavity prole [1]. Many commercial solidication simulation softwares can
be used to obtain reliable simulation results if the appropriate data
of thermal properties and boundary conditions are provided [2].
For the heat transfer in solidication, how the heat transfers
through the castingmould interface is one of the most important
boundary conditions to be characterized because this problem directly dominates the evolution of solidication and controls the
freezing conditions within the casting. Therefore, the determination of interfacial heat transfer coefcient (IHTC) is vital ahead of
the simulation of the solidication process. In fact, the IHTC depends upon multiple factors such as die coating thickness, insulating pads, chill and casting geometries, pouring temperature,
surface roughness, alloy composition, metallostatic head, mould
* Corresponding author. Tel./fax: +86 29 88495774.
E-mail address: zhangwh@nwpu.edu.cn (W. Zhang).
0894-1777/$ - see front matter 2010 Elsevier Inc. All rights reserved.
doi:10.1016/j.expthermusci.2010.03.009

temperature and other mechanical boundary conditions [36]. Its


determination is often carried out by manual adjustments to reduce the difference between the experimental observation and
the numerical prediction.
Generally, two kinds of methods exist. The rst one is to measure the size variation of the interfacial gap that usually appears
at the metal/mould interface during the solidication process. For
example, Prates and Biloni [7] and Nishida et al. [8] measured the
IHTCs based on the immersion method, uidity test, unidirectional
method and one-dimensional solidication in a mould. The formation process of the air gap and the involved heat transfer mechanism were investigated by measuring the displacements and
temperatures for both cylindrical and at castings of aluminum alloys. The second one is to evaluate the IHTC inversely based on the
temperature data measured at selected locations in both the casting
and the mould or chill. Note that the surface temperature or heat
ux is determined based on the measured temperatures at internal
points near the surface. Since the solidication of a casting involves
both the material phase change and the variation of thermal properties with respect to the temperature, the inverse heat conduction
is a nonlinear problem and can be solved by means of the nonlinear
estimation methods [9,10]. For instance, Lau et al. [11] studied the
IHTC between an iron casting and a metallic mould. Souza et al. [12]
analyzed the heat transfer along the circumference of cylinders
made up of SnPb alloys in the mould.

1069

W. Zhang et al. / Experimental Thermal and Fluid Science 34 (2010) 10681076

Nomenclature
a1, a2
b1, b2
fs
h
k
L
m
n
q
t
tc
T

Subscripts
0
initial state
c
casting
cr
critical
h
heat transfer coefcient
l
liquidus
m
mould
s
solidus
T
temperature

coefcient of castingmould IHTC function


coefcient of mouldenvironment IHTC function
solid fraction
heat transfer coefcient
number of thermocouples in the mould
latent heat of fusion
number of time steps
number of thermocouples in the casting
interface heat ux
solidication time
critical solidication time
temperature

Superscripts
est
predicted values
exp
experimental data
max
maximum
min
minimum

Greek symbol
k
thermal conductivity

However, few reliable data of the IHTCs are available for the
investment casting process in practice. Sturm and Kallien [13]
identied the IHTC involved in the model of an aluminum alloy
investment casting where the resultant data of IHTC (1000 W/
m2 K) was assumed to be unchanged throughout the solidication.
Anderson et al. [14] combined the simulation and experiment to
study thermal behaviors of a two-dimensional symmetrical aluminum casting where the IHTC was buried in an overall heat transfer
coefcient. Based on the nonlinear estimation technique mentioned above, Sahai and Overfelt [15] completed a study of the
IHTC for both cylindrical and plate investment castings of a
nickel-based alloy. For the cylindrical casting (mould preheated
to 745 C), it was found that the IHTC varied linearly from
200 W/m2 K at 1300 C to 100 W/m2 K at 850 C. For the plate casting, the IHTC was found to vary between 5000 W/m2 K at 1400 C
and 100 W/m2 K at 1100 C. The results showed that the casting
shape had a great impact upon the IHTC in the investment casting.
OMahoney and Browne [16] suggested that cares should be taken
of the solidication process, the alloy type and the metallostatic
head effect. The aluminum casting alloys, 413, A356, 319, were
used in their study.
For these reasons, this work is to develop a simple and universal
inverse methodology, which makes use of the existing simulation
softwares such as ProCAST to resolve the IHTCs in the investment
casting process of a complicated blade. Based on a switch function
of solidication time, a novel model of IHTC is proposed to replace
the original power function. With the obtained IHTCs, the predicted temperature is compared with the experimental data. Besides, thermocouples are placed in a very thin mould cavity
without manufacturing a special mould. This methodology is helpful for a foundry engineer to look for a reference effectively on how
to apply boundary conditions for simulation of a specic casting
process.
2. Mathematical model of casting process
Fig. 1 depicts the heat transfer through between the two contacting surfaces. When the mould is suddenly lled with the liquid
metal, the effects of uid ow in the liquid phase, the convective
heat transfer and the radiative heat transfer are negligible. Therefore, the direct problem for the casting region is formulated only
in terms of unsteady-state heat conduction.







@T
@
@T
@
@T
@
@T
@fs

qL
k
k
k
qc
@t @x
@x
@y
@y
@z
@z
@t

q = h (Tc Tm )

Casting

Tc

Mould

Casting

Mould

Tm

Insulating heat material


Fig. 1. Schematic of a castingmould interface.

where q is the cast density, c and k are specic heat and thermal
conductivity, respectively. L is the latent heat of fusion and fs is
the solid fraction. Note that the thermal properties are known during the investment process. The initial and boundary conditions for
the casting region are

initial condition

Tjtt0 T 0 x; y; z

at castmould interface  k

@T
q hc T  T m
@n

2a
2b

The casting temperature eld is governed by the above heat conduction equation and boundary conditions. Numerical solutions
can be obtained by means of the nite element method.
Obviously, hc, the IHTC at the castingmould interface, affects
the calculated temperature eld and is thus of importance for
the numerical solution of the casting temperature. Likewise, the
governing equation related to the mould region is similar to the
above one except that the source term, qL @f@ts , is not included. Moreover, hm at the mouldenvironment interface has to be determined
in advance. For an inverse heat transfer problem, the aim is to predict the unknown IHTCs from the knowledge of measured or/and
calculated temperatures at specic positions on the interface. This
paper is to determine hc and hm in the blade investment casting
process.
3. Determination of interfacial heat transfer coefcient
3.1. Inverse parameter estimation
Inverse estimation methods are based on the minimization of
an objective function containing both estimated and measured

1070

W. Zhang et al. / Experimental Thermal and Fluid Science 34 (2010) 10681076

1.0

temperatures. Solving one such optimization problem without


constraints can identify unknown parameters involved in the
model. In this study, a globally convergent method (GCM), originally proposed by Svanberg [17], is used to nd the optimal values
of unknown parameters by minimizing an objective function dened by

w2

2
Pm Pk  exp
 T est
ij
i1
j1 T ij
k

est
@T est
T est
ij h
ij h1 ; . . . ; hr dhr ; . . . hn  T ij h1 ; . . . ; hr ; . . . hn

@hr
dhr

Then an iterative procedure is designed to nd the minimization solution of S(h). It must be pointed out that nite difference
method used for the sensitivity analysis suffers from two major
drawbacks. Firstly, the approximation accuracy depends on the
magnitude of the perturbation dhr. If dhr is too small, the roundoff errors will be signicant. Oppositely, if dhr is too large, the
truncation errors will degrade the accuracy. In this work, the
magnitude of perturbations is automatically chosen by an optimization method. Secondly, the use of nite difference method is
expensive because the nite element reanalysis must be run
n + 1 times for each iteration. At this point, an efcient way of
decreasing the computing cost is to parameterize the IHTC only
as a function of time because of the interdependence between
the IHTC and the temperature.
3.2. Continuous IHTC model with three-step evolution
To achieve a reasonable model of the castingmould IHTC, the
complicated mechanism of the interfacial heat transfer in the progressive solidication should be discussed rstly. In general, the
variation of the castingmould IHTC with time can be divided into
four stages: (i) At the rst stage, the IHTC increases rapidly when
the molten alloy is poured into the mould. Although the ow in
the alloy has a great inuence on the IHTC, it is not considered in
this study due to the limitation of high frequency acquisition disposals. (ii) At the second stage, the IHTC is higher in longer mushy
zones with the temperature variations between liquidus temperature and solidus temperature, as pointed out by Santos et al. [18].
The magnitude of the IHTC almost remains unchanged because the
macro air gap does not appear during such a short period of time.
(iii) At the third stage, the IHTC starts to decrease rapidly as long
as the casting thickness becomes larger and larger with a decrease
of the velocity of heat transfer from the casting to the mould. (iv)
At the fourth stage, a gradual decrease of the IHTC is observed
due to the further increase of the air gap.
Based on the above interface heat transfer mechanism, a new
model of the IHTC is proposed with the negligence of the rst
stage. The IHTC could be assumed to be a constant at the second
stage, whose initial value may change from case to case. A power
function of time is used to characterize the signicant drop of

tcr=10

=10

0.6

0.4
=100

3
exp
where T est
denote the estimated and the experimental data
ij and T ij
of the temperature eld at various thermocouple locations and time
increments, respectively. m is the number of time steps, n is the
number of thermocouples in the casting, and k is the number of
thermocouples in the shell mould. w1 and w2 are the weightings.
Because the minimization must ensure the accuracy of the temperature eld over the casting as much as possible, the value of w1 is
often larger than that of w2. Here, w1 and w2 are assumed to be
0.7 and 0.3, respectively.
In order to minimize S(h), the rst-order sensitivity coefcients
are usually calculated by nite difference scheme with

Switch function

=3

-1

1+e t-tcr

Sh w1

2
Pm Pn  exp
 T est
ij
i1
j1 T ij

=1

0.8

0.2

0.0
0

10

15

20

t
Fig. 2. Typical switch functions for heat transfer coefcient.

the IHTC caused by the appearance of the macroscopic air gap during the third stage and the fourth stage. Therefore, a piecewise
function is thus proposed to formulate the castingmould IHTC.


h

h0

t  tcr

a1 t a2

t > tcr

where h0 is an initial value of the IHTC in the initial stage; a1 and a2


are the parameters to be determined. tcr is the critical time corresponding to the intersection between the second stage and the third
stage.
In order to improve the prediction and optimization efciency, a
continuous and differentiable switch function is devised to express
the heat transfer coefcient. The castingmould IHTC is now reformed as

hc h0

1
1
a1 ta2
1 eattcr
1 eatcr t

where a refers to a large positive number. The term, 1ea1ttcr , denotes a typical switch function, as illustrated in Fig. 2 for different
values of a when tcr = 10. Clearly, a large a results in a closed
approximation of the unity once t is less than tcr, or of the zero when
t is larger than tcr. Thus, a moderate value of a = 10 is chosen in the
model of the castingmould IHTC and the three unknown parameters, a1, a2 and tcr, are to be determined.
As to hm, the external surface temperature of the shell mould is
initially low. It rises rapidly to a peak value at the beginning of
solidication and then declines. According to the experimental
data [18], the values of the IHTC are 22 W/(m2 K) and 34 W/
(m2 K) when the temperature of the mould surface is 300 C and
600 C, respectively. Similarly, a power function is given to correlate the IHTC with the process time

hm b1 tb2

where b1 and b2 are the unknown parameters to be determined.


3.3. Description of the optimization procedure
The optimization procedure is shown in Fig. 3, where the GCM
is used as the optimizer. The process starts by initializing the basic
data for the direct heat transfer analysis and the optimization programs. Based on initially estimated parameters and sensitivity values, a proper search direction and a step size will be evaluated by
the optimizer to update design parameters. In this study, ve
parameters tcr, a1, a2, b1 and b2 are optimized. Because the objective function is highly nonlinear, the nite difference method is
applied in sensitivity analysis.

W. Zhang et al. / Experimental Thermal and Fluid Science 34 (2010) 10681076

1071

DATA INPUT
Initialize design parameter & their upper and lower bounds

FEM ANALYSIS
Call ProCAST to simulate solidification
and calculate objective function

MODIFY INPUT
Refresh the design parameter by GCM

SENSITIVITY ANALYSIS
Calculate the sensitivity of the design parameter

No
Converged ?

Yes
Stop

Fig. 3. The owchart of optimization procedure.

Optimization problems are usually written in the following


form:

Min f0 x
s:t:

fi x  0

i 1; . . . ; m

Support

x2X

f0, f1, . . . , fm are real valued functions which are assumednto be second-order continuously differentiable on the set X x 2 Rjxmin
j
; j 1; . . . ; ng.
 x  xmax
j
To avoid the situation that the feasible design space is empty, a
modied optimization problem is considered:

Min f0 x

m 
P
i1

s:t:

Mi yi y2i =2

fi x  yi  0


9
i 1; . . . ; m

where Mi are often assigned by very large real numbers and yi are
so-called articial variables. All yi are usually zeros at the optimum
unless some of them are relaxed to take positive values.
The GCM works iteratively according to the following general
scheme: Assume x represents the set of design parameters. During
iteration k, an explicit approximate sub-problem is generated at a
current iteration point (x(k), y(k)). In the sub-problem, the functions
fi(x) are replaced by approximate convex functions based on the
gradient information and the information from the previous iteration points. Once this sub-problem is solved, the optimal solution
becomes the starting point of the next iteration for the new subproblem. A description of the GCM can be found in [17].
4. Experimental setup of the investment casting process of a
blade
To validate the proposed IHTCs models, aluminum alloy A355 is
used instead of super-alloy in the present work. Moreover, to reduce experimental cost, the gravity casting process is adopted.
The procedure starts with a blade fabricated by an investment
casting wax (the pattern). The wax is heated above its melting temperature and then pressed into a steel die. The wax pattern is made
up of two parts: the core and the exterior, as shown in Fig. 4.

Fig. 4. The core and the exterior of the wax pattern.

Six thermocouples are positioned in the patterns middle crosssection as shown in Fig. 5. Six wood sticks are selected to drill some
holes of 1.5 mm in diameter. Then these sticks are inserted into six
holes of the wax blade that is xed to the feed system, as shown in
Fig. 6. Finally the pattern is cleaned to allow the adherence of the
mould material. The investment shell moulds are composed of two
layers. Firstly the pattern is dipped into the ceramic slurry and
drained, and then rained by ne ceramic and nally dried in a
vent-pipe. This procedure is repeated until a desired thickness of
2.0 mm attains for the primary shell. The other six sticks are used
to measure the temperature of the primary shell. A secondary layer
with a thickness of 4.0 mm is formed in the same way. When the
wood sticks are burned out, 12 K-type thermocouples are then
placed into the small holes with a depth of about 2.0 mm from
the interface to metal region and from the interface to the mould
region, as shown in Fig. 5. Moreover, two thermocouples are placed
on the external surface of the shell mould so as to acquire the

1072

W. Zhang et al. / Experimental Thermal and Fluid Science 34 (2010) 10681076

The middle cross-section


thermocouple
Feed System
0
2.

.0

4
2
8

The middle
cross-section

60

110

ch
or
d

sand_zircon

Casting
Blade

ax
ial

11

200

casting

10

support

sand_silica2

thermocouple

A-A
6
12

Fig. 6. The 3D model of the blade and the feeder.

The interface

Fig. 5. The cross-section of the blade and thermocouple positions.

temperature data for the prediction of the IHTC between the mould
and the environment.
The mould shell is not preheated in gas furnace. Molten aluminum alloy is poured into the mould shell at a temperature of about
624 C by the gravity method, and the mould shell is cooled by the
air with the insulating heat materials on the top and the bottom, so
that the heat ux from the casting to the mould shell can only take
place along the periphery of the turbine blade cross-section as
shown in Fig. 7. The temperature is recorded by sampling frequency
of 1 Hz using a temperature instrument HR3200 (YOKOGAWA,

Japan). Before the thermal proles are measured, they must be


smoothed out using a digital lter.
Although the experimental setup should be designed to be as
representative as possible of the real process, one should realize
in mind that it is impossible to consider performing an inverse calculation on a real 3D casting geometry due to the prohibitive computing time. Thus, a 2D geometry is considered in the parameter
optimization process. In addition, a certain number of thermocouples have to be properly located rather than ooding the experiments with many thermocouples. With the experimental setup,
the maximum absolute error is about 12 C.
The relevant properties and chemical composition of aluminum
alloys are widely available in the literature [19], but relatively little
accurate information of the investment of the casting and shell

Heat insulating material

The Shell
mould

The Casting
Atmosphere

Heat insulating material


Fig. 7. Schematic representation of the experimental setup.

1073

W. Zhang et al. / Experimental Thermal and Fluid Science 34 (2010) 10681076

650

Table 1
Thermal data for aluminum alloy A355.

A355

Liquidus
temperature (C)

Eutectic
temperature (C)

Solidus
temperature (C)

Freezing
range (C)

624

582

540

84

Table 2
Thermal data for the shell mould.
Shell
mould

Conductivity
(W/(m K))

Density
(kg/m3)

Special heat
(kJ/kg/K)

Sand silica
Sand zircon

0.59
0.83

1520
2780

1.20
0.77

600

Experiment thermocouple 1,7


Simulated

550

Temperature (oC)

Alloy

The casting

500
450
400
350
300
The shell mold

250

200

1400

180

1200

Ethalphy

100

200

300

400

500

Time (s)
Fig. 9. Experimental and predicted temperature at the external surface of the shell
mould.

1000
800

140
600
120

400

vity

ucti

100

d
Con

200

80

0
0

200

400

600

800

Temperature (oC)
Fig. 8. Conductivity and enthalpy of aluminum alloy A355.

300

Mould-Enviroment HTC (W/(m2.K))

160

Ethalphy (kJ/kg)

Conductivity (W/(m.K))

200

250
200

Mould-enviroment heat transfer coeffecient

hm=82.06 t-0.26

150
100
50
0
0

mould is available. In this paper, the thermal data of the shell


mould is referred from the thermo-physical properties given in
the database of the ProCAST and the literature [20]. Thermo-physical properties of the casting and the shell mould are given in
Tables 1 and 2 and Fig. 8.

100

200

300

400

Time (s)
Fig. 10. Variation of the mouldenvironment IHTC of alloy A355.

650
Tl = 624

5. Results and discussion

Fig. 9 shows the temperature variations of thermocouples 1 and


7 sampled in the primary layer of the casting and shell mould (as
shown in Fig. 5) during the solidication experiment, respectively.
Since the heat transferred between the shell mould and the environment includes convective and radiative heat, it can be observed
that the temperature on the external surface of the shell mould
rises rapidly from the beginning of solidication to a peak value
and then declines. With the help of the power function introduced
in Eq. (7), a docent agreement is achieved between the measured
temperatures and the numerical predictions, as shown in Fig. 9.
In this case, the mouldenvironment IHTC by the optimization process is presented in Fig. 10.

600

Thermocouple 1,3,5
Ts = 582

Temperature (oC)

5.1. The mouldenvironment IHTC

Thermocouple 2,4

550

500
Thermocouple 6

450

400
0

tc = 58s

50

100 150 200 250 300 350 400 450 500

Time (s)
Fig. 11. Experimental temperature proles of the thermocouples.

5.2. The metal castingmould IHTC


Considering the geometrical feature of the blade as shown in
Fig. 11, the cross-section of the blade can be divided into four
parts: the trailing edge (thermocouple 6), the leading edge (thermocouple 3), the concave side (thermocouples 1 and 2) and the

convex side (thermocouples 4 and 5). The slopes of different temperature curves represent the cooling rates at the corresponding
measured positions. In general, the cooling curve of the casting
consists of three stages: beyond the liquidus, between the liquidus
and the solidus, and below the solidus. At the rst and third stages

1074

W. Zhang et al. / Experimental Thermal and Fluid Science 34 (2010) 10681076

Table 3
Optimization results by GCM.
Model

Lewis and Ransing [21]

Parameters
Values

C
7100.21

Santos et al. [18]

n
0.21

a1
13.40

a2
1189.94

This study
2

a3
0.53

h0 (W/m K)
12160.36

a2
0.52

b1
82.06

b2
0.28

90

650

80

This work
By Lewis et al. [21]
By Santos et al. [18]
Experimental data

600

Lewis et al. [21]


Santos et al. [18]
This work

70
60

550
500

Error( )

Temperature (oC)

a1
1245.61

By L
ewis

50
40

average error=33.27

30

450

average error=20.00

20
By S
a

400
0

100

200

300

ntos

400

average error =7.85

10
0

500

100

200

Time (s)
Fig. 12. Comparisons between the experimental temperature and the predicted
temperature by different models at thermocouple 1.

300

400

500

Time (s)
Fig. 13. Absolute errors of predicted temperatures with different models.

Temperature ( )

660
640

Thermocouple 6 simulated

620

Thermocouple 3 simulated
Thermocouple 3 experiment
Thermocouple 6 experiment

600
580
560
540
520
500

10

20

30

40

50

60

70

80

90

80

90

Time (s)

(a) Thermocouples 3 and 6.


640
Thermocouple 2 simulated
Thermocouple 4 simulated
Thermocouple 2 experiment
Thermocouple 4 experiment

620

Temperature ( )

the temperature of the casting falls rapidly, while at the second


stage the cooling rate decreases slowly because of the release of
the latent heat. However, both the experiment data and those data
published in the literature [2,20] showed that due to the high temperature difference between the casting and the shell mould, the
temperature of the casting surface decreased so quickly that the
second stage almost disappeared. Once the temperature falls below the solidus, the effect of IHTC on the cooling rate is much
stronger because the heat conduction between the casting and
the shell mould dominates the cooling rate of the casting.
Note that the cooling rate is the fastest in the trailing edge
where the blade prole is the thinnest. Therefore, the shrinkage
happens earlier in the trailing edge than in other positions and
thereby the macro gap forms earlier in this position. Accordingly,
the slope of its cooling curve changes more early. The cooling rate
of the concave side is slower than that of the convex side due to the
heat radiation. The rapid decrease of these cooling prole slopes at
about 58 s indicates that the macro airgap forms at this critical
time rather than at the solidication time of the interface. Such
critical time is later than the solidication time of the interface.
Therefore, the macro airgap will not be formed until the interfacial metal skin between the casting and the shell mould becomes
effective to resist the action of the metallostatic head from liquid
metal. As all cooling rates are not so distinct, a unique equivalent
interfacial heat transfer coefcient is used for simulation. According to the above analysis, the value of h is constant for t < 58 s
and then modeled by a power expression of time for t < 58 s. Based
on this feature, Eq. (6) is used to resolve the IHTC in this study.
The other two models proposed by Santos et al. [18] and Lewis
and Ransing [21] are also tested to calculate the IHTC, and the optimization results are given in Table 3. The corresponding temperature calculated at thermocouple 1 is shown in Fig. 12. Differences
between the predicted results and the measured data are shown
in Fig. 13. Among the three IHTC models, the model proposed in
this study is found to provide the best tting to the experimental
data with an average error of about 7.85 C, while the other two
models produce the average errors of about 33 C and 20 C,
respectively.

600
580
560
540
520

10

20

30

40

50

60

70

Time (s)

(b) Thermocouples 2 and 4.


Fig. 14. Comparisons between the experimental temperatures and the predicted
temperatures.

24.
28s

22. 2
8s

W. Zhang et al. / Experimental Thermal and Fluid Science 34 (2010) 10681076

the interface decreases rapidly due to the increase of the transient


thermal resistance between the metal and the mould. Consequently, the IHTC reaches a relatively lower value of about 50 W/
m2 K, as shown in Fig. 16.
Furthermore, based on numerical tests, the same convergent
solution of the objective function is achieved even if the GCM
method starts with different initial values of design parameters.
The computational cost is decreased signicantly owing to the
parameterization of the IHTC, which reduces the number of design
parameters effectively.

ax

Rm

8s
.2
26
.2
24

8s

8s
.2
22

The inscribed circle

.
18

6. Conclusions

s
28
.
14

Solidus

s
28
8s
.2
10
6.

s
28

Fig. 15. Variations of solidus isotherms as a function of time.

400

Casting-Mould IHTC (W/(m2.K))

1075

The interfacial heat transfer coefcients (IHTCs) in the investment casting of a solid blade have been investigated on the basis
of an experimental study and an optimization method. A commercial software ProCAST and an optimization tool with globally convergent method (GCM) are employed.
Equivalent parameterized models of the IHTCs including a continuous three-step evolution for the castingmould IHTC and a
power function of time for the mouldenvironment IHTC are proposed. Involved parameters in the model are resolved by the GCM
optimization method. Good agreements between the experimental
and the predicted temperatures are achieved with the maximum
deviation being less than 5.5%. Even with different starting conditions of design parameters, the convergence can be achieved
efciently.

350

Acknowledgement

300

Casting-Mould heat transfer coeffecient

250

This work was supported by a grant from National Science Fund


for Distinguished Young Scholars (No. 10925212).

hc=1245.61 t-0.52

200
150

References

100

[1] S.C. Modukuru, N. Ramakrishnan, A.M. Sriramamurthy, Determination of the


die prole for the investment casting of aerofoil-shaped turbine blades using
the nite element method, J. Mater. Process. Technol. 58 (1996) 223226.
[2] J.M. Drezet, M. Rappaz, G.U. Grm, M. Gremaud, Determination of
thermophysical properties and boundary conditions of direct chillcast
aluminum alloys using inverse methods, Metall. Mater. Trans. A 31A (2000)
16271634.
[3] H.M. S
ahin, K. Kocatepe, R. Kaykc, N. Akar, Determination of unidirectional
heat transfer coefcient during unsteady-state solidication at metal casting
chill interface, Energy Convers. Manage. 47 (2006) 1934.
[4] T.A. Blase, Z.X. Guo, Z. Shi, K. Long, W.G. Hopkins, A 3D conjugate heat transfer
model for continuous wire casting, Mater. Sci. Eng. A 365 (2004) 318324.
[5] M.A. Gafur, M.N. Haque, K.N. Prabhu, Effect of chill thickness and superheat on
casting/chill interfacial heat transfer during solidication of commercially pure
aluminum, J. Mater. Process. Technol. 133 (2003) 257265.
[6] M.A. Martorano, J.D.T. Capocchi, Heat transfer coefcient at the metalmould
interface in the unidirectional solidication of Cu8%Sn alloys, Int. J. Heat Mass
Trans. 43 (2000) 25412552.
[7] M. Prates, H. Biloni, Variables affecting the nature of the chill zone, Metall.
Trans. 3 (1972) 15011510.
[8] Y. Nishida, W. Droste, S. Engler, The airgap formation process at the casting
mold interface and the heat transfer mechanism through the gap, Metall.
Trans. B 17B (1986) 833844.
[9] J.V. Beck, Determination of optimum transient experiments for thermal
contact conductance, Int. J. Heat Mass Trans. 12 (1969) 2133.
[10] J.V. Beck, Nonlinear estimation applied to the nonlinear inverse heat
conduction problem, Int. J. Heat Mass Trans. 13 (1970) 703716.
[11] F. Lau, W.B. Lee, S.M. Xiong, B.C. Liu, A study of the interfacial heat transfer
between an iron casting and a metallic mould, J. Mater. Process. Technol. 79
(1998) 2529.
[12] E.N. Souza, N. Cheung, C.A. Santos, A. Garcia, Factors affecting solidication
thermal variables along the cross-section of horizontal cylindrical ingots,
Mater. Sci. Eng. A 397 (2005) 239248.
[13] J.C. Sturm, L. Kallien, Solidication simulation of an integrated aircraft
structural component, in: Modelling of Casting, Welding and Advanced
Solidication Processes V, 1990, pp. 2330.
[14] J.T. Anderson, D.T. Gethin, R.W. Lewis, Experimental investigation and
numerical simulation in investment casting, Int. J. Cast Metals Res. 9 (1997)
285293.

50
0

100

200

300

400

Time (s)
Fig. 16. The castingmould IHTC of alloy A355 at last two stages.

Comparisons between the experimental temperature and the


predicted temperature in other locations are shown in Fig. 14. It
is observed that an acceptable agreement is achieved between
the experimental data and those predicted results. The maximum
deviation is less than 5.5%. Solidus isotherm varying as a function
of time is shown in Fig. 15. It can be seen that the solidication
ends around the center of the maximal inscribed circle of the blade
prole. This indicates the shrinkage center of the blade.
From optimization results in Table 3, it is found that the IHTC
takes a very high value, i.e., up to 12160.36 W/m2 K, at the initial
stage of solidication as a result of the tight surface contact between the liquid metal and the shell mould. The shell temperature
rises rapidly from the beginning of solidication, since the shell
mould is very thick and is not preheated. As a result, the mould
expansion favors the thermal contact between the metal and the
shell surface. Therefore, the initial value of the IHTC in the investment process is higher. Moreover, it might be deduced that the initial value increases with increasing values of superheat, and the
rst stage in which the value keeps a constant will be prolonged.
Once the air gap has been formed, the heat transferring across

1076

W. Zhang et al. / Experimental Thermal and Fluid Science 34 (2010) 10681076

[15] V. Sahai, R.A. Overfelt, Contact conductance simulation for alloy 718
investment castings of various geometries, Trans. Am. Foundrymens Soc.
103 (1995) 627632.
[16] D. OMahoney, D.J. Browne, Use of experiment and an inverse method to study
interface heat transfer during solidication in the investment casting process,
Exp. Therm. Fluid Sci. 22 (2000) 111122.
[17] K. Svanberg, A globally convergent version of MMA without line-search, in:
Proceedings of the First World Congress of Structural and Multidisciplinary
Optimization, Goslar, Pergamon, Germany, 1995, pp. 916.

[18] C.A. Santos, J.M.V. Quaresma, A. Garcia, Determination of transient interfacial heat
transfer coefcients in chill mold castings, J. Alloy. Compd. 319 (2001) 174186.
[19] Foundry Manual. China Machine Press, 1993. 2 (in Chinese).
[20] H.L. Zeng, The Interfacial Heat Transfer Behavior between High Temperature
Alloys and Ceramic, Master thesis. National Cheng Kung University, Taiwan
40-5, 2002.
[21] R.W. Lewis, R.S. Ransing, A correlation to describe interfacial heat transfer
during solidication simulation and its use in the optimal feeding design of
castings, Metall. Mater. Trans. B 29B (1998) 437448.

You might also like