Sie sind auf Seite 1von 6

Journal of Materials Processing Technology 211 (2011) 181186

Contents lists available at ScienceDirect

Journal of Materials Processing Technology


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

Determination of heat transfer coefcient and ceramic mold material


parameters for alloy IN738LC investment castings
C.H. Konrad, M. Brunner, K. Kyrgyzbaev, R. Vlkl, U. Glatzel
Metals and Alloys, University Bayreuth, Ludwig-Thoma-Str. 36b, D-95447 Bayreuth, Germany

a r t i c l e

i n f o

Article history:
Received 22 April 2010
Received in revised form 27 August 2010
Accepted 31 August 2010

Keywords:
Investment casting
Heat transfer coefcient
IN738LC
Superalloys

a b s t r a c t
Investment casting molds with different numbers of shells and pre-heating temperatures were investigated in this study. The primary layer consists of colloidal silica bound ZrSiO4 with additions of CoAl2 O4
to achieve ne grains and to reach a good surface quality, whereas the following layers consist of
mullite bound by colloidal silica. Interface temperatures (alloy/mold) that are necessary to determine
heat transfer coefcients were obtained by linear extrapolation. Heat transfer coefcients in the range
of 300660 W/(m2 K) were obtained. The castings were examined with regard to grain size and secondary dendrite arm spacing. Physical properties of the investment casting mold were examined by
differential scanning calorimetry (DSC) and Laserash methods for temperatures up to 1300 C. The specic heat capacity was determined to 1.13 J/(g K), thermal diffusivity was found to be in the range of
(45) 107 m2 /s and the thermal conductivity to be 1 0.1 W/(m K).
2010 Elsevier B.V. All rights reserved.

1. Introduction
Investment casting is a well established manufacturing process
for blades and vanes in aircraft engines and stationary gas turbines.
These parts have to fulll very strict quality requirements in order
to withstand high mechanical loads at temperatures up to 1050 C
for several thousand hours. The alloys used for these purposes are
mainly nickel based superalloys such as the alloy Inconel 738 low
carbon (IN738LC), which serves as model alloy in this work.
Simulation of the investment casting process is expected to
shorten development times in industry. However, many material
parameters of the alloy, the mold and the alloy/mold interface have
to be known accurately.
Liu et al. (1998) measured the liquidus and solidus temperature
for the alloy IN738LC and showed the interdependency of the melt
temperature with grain size. Chapman et al. (2008) measured physical properties of this alloy and of different ceramic mold materials.
An important parameter is the heat transfer coefcient (HTC)
which describes the temperature drop in the contact zone of
the melt and the mold during solidication. Ozisik et al. (1993)
described an inverse method to solve the heat conduction equation in order to determine the heat transfer coefcient. OMahoney
and Browne (2002) recorded carefully the transient temperatures inside ceramic investment casting shell moulds and inside
aluminum melts during solidication. With this data they could

Corresponding author. Tel.: +49 921 55 5555; fax: +49 921 55 5561.
E-mail address: uwe.glatzel@uni-bayreuth.de (U. Glatzel).
0924-0136/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.jmatprotec.2010.08.031

numerically solve the inverse heat conduction problem for the


unknown heat transfer coefcient by nite differences. Sahai
and Overfelt (1995) too solved inverse problems for plate and
axisymmetric geometries with the commercial software ProCast
for investment casting of the nickel based alloy 718. Aweda and
Adeyemi (2009) determined the heat transfer coefcient between
steel molds and aluminum alloys during squeeze casting. The
transient interface temperatures versus time were identied by
polynomial curves tting. They could verify their approach with
the results from Krishnan and Sharma (1996).
In this work a method is presented to deduce the heat transfer
coefcient between alloy and investment casting mould directly
from recorded transient temperature data.
2. Experimental
2.1. Investment casting
All experiments were carried in a proprietary vacuum investment casting furnace (Fig. 1) with a minimal pressure of
3 105 mbar. Two 40 kW medium frequency induction heating
systems with separate inductor coils are installed. The upper inductor is used for melting the master alloy (IN738LC) in an alumina
crucible at temperatures up to 1500 C. The second induction heating system preheats the ceramic shell mold through a graphite
susceptor. When the mold has reached a predened temperature
it is lifted into the pouring position in a few seconds. The whole
casting process, i.e. the preheating of the mold, the melting of the
master alloy and the pouring can be controlled and monitored with

182

C.H. Konrad et al. / Journal of Materials Processing Technology 211 (2011) 181186
Table 1
Mold-parameters for the heat transfer coefcient experiments.
Number of layers

Mold preheat temperature

8
6
4

900 C

1000 C

1100 C

1235 C

T900L6

T1000L8
T1000L6
T1000L4

T1100L6

T1235L6

in all tests and was measured by an insulated thermocouple


type S.
2.2.3. Mold parameters
Specic heat capacity cp of the ceramic mold were measured
with a Linseis L81/1550 DSC (Selb, Germany) in the temperature
interval 2001200 C. Cross-checks were performed with a Netzsch 404 Pegasus DSC (Selb, Germany) in the temperature interval
2001300 C. Both measurement series were carried out under
argon atmosphere and repeated three times with different samples.
Thermal diffusivity a of the mold material was measured with a
Netzsch LFA427 laser ash. Two measurements were undertaken.
A round ceramic sample (12.7 mm radius and 2.04 mm thickness)
was used in the temperature interval 9001300 C and a square
ceramic sample (10 mm 10 mm and 1.97 mm thickness) was used
in between 15 and 1300 C. Thermal conductivity  was derived
from thermal diffusivity measurements according to Eq. (1):
(T ) = a(T )  cp (T )

Fig. 1. Schematic drawing of the vacuum investment casting furnace.

16 thermocouples simultaneously. In addition, pouring of the melt


and subsequent cooling of the mold can be monitored with an
infrared-camera through a ZnSe inspection window. Typical ingot
mass is about 300 g of nickel based alloy. Cross-sections of the
cast ingots were examined to measure grain size and secondary
dendrite arm spacing.
2.2. Materials
2.2.1. Shell mold
The shell mold with a feeder system was manufactured using
a conventional lost-wax process (see Fig. 2). The chosen plate
geometry (60 mm 40 mm 4 mm) allows the assumption of a
one-dimensional heat transfer problem in the direction of the plate
normal. Molds with 4, 6 and 8 layers and different preheat temperatures were tested. Due to the different numbers of layers the
thickness of the molds varies from 4 to 11 mm. All layers besides
the innermost layer consist of mullite particles of different sizes
bound by colloidal silica. The innermost layer consists of colloidal
silica bound ZrSiO4 with additions of CoAl2 O4 for grain renement
and a good surface quality. The parameters of the casting tests,
denominated as T1000L4, T1000L6, T1000L8, T900L6, T1100L6 and
T1235L6 are given in Table 1.
2.2.2. Alloy
The alloy IN738LC serves as model alloy throughout this
work. The nominal composition of IN738LC is given in Table 2.
It is an alloy frequently used for polycrystalline and directionally solidied (Reed, 2006) lost wax investment casting of
turbine blades. The pouring temperature of the melt was 1500 C

(1)

The density  of ceramic mold was deduced by dividing


the mass by the volume at room temperature. The density was
1.80(0.3) g/cm3 . Because density of the ceramic shell mold is supposed to change only slightly over the interesting temperatures its
temperature dependence was not accounted for.
2.3. Instrumentation
The molds were instrumented with 4 thermocouples (Fig. 2).
A ceramic insulated thermocouple type S with wire diameters
of 0.25 mm and sealed by sintered alumina slurry was used to
measure the temperature of the melt after pouring. Three thermocouples type K with a diameter of 0.12 mm were used to
measure the transient temperatures inside the ceramic shell mold.
Two were placed in different depths of the mould. The third one
was used to measure the surface temperature of the mold. All
thermocouples were placed near the center of the mold, where
the heat transfer is assumed to be one-dimensional, i.e. the heat
is almost entirely transferred perpendicular to the mold wall.
The thermocouples were xed in drilled holes with diameters of
about 1 mm with high temperature resistant glue (Ultra-Temp
516, Kager, Germany) which consists of sodium silicate bound
SiO2 and ZrO2 . Weathers et al. (2006) found that the ideal way
of placing thermocouples in ceramic molds is parallel to the
isotherms. Therefore the drill holes were inclined by 45 normal
to the isotherms in order to keep the inuence of the thermocouples on the heat propagation as low as possible (see Fig. 2).
The exact distances from the heads of the thermocouples to the
interface were measured after the mold was removed from the
casting.

Table 2
Nominal composition of IN738LC (Reed, 2006).
Element

Ni

Cr

Co

Mo

Al

Ti

Ta

Nb

Zr

Weight-%

Bal.

16

8.5

1.75

3.4

3.4

1.75

0.9

0.11

0.01

0.04

C.H. Konrad et al. / Journal of Materials Processing Technology 211 (2011) 181186

183

Fig. 2. a) Wax model; b) mold instrumentation (cross-section).

3. Theory
The heat transfer coefcient h of the alloy/mold interface
depends on the total heat ux q through the interface and the
temperature drop T at the interface:
h=

q
q c + q s
=
T
T

(2)

The total heat ux consists of the ux q c the melt looses to the


mold and the ux q s due to the exothermic nature of the solidication. The ux q c is calculated by the temperature decrease in the
considered volume:
q c =

s x cp (T (t) T (t + t))
t

For the calculation, several simplications have been made. The


heat transfer in the alloy is much faster than in the mold, hence
the alloy is assumed to have everywhere the temperature measured by the thermocouple in the center of the casting at position
1 mm in Fig. 3. The temperature at the mold side of the interface
is estimated by linear extrapolation of the temperatures measured
by the thermocouples in the mould. The difference between these
two temperatures at the interface is the temperature drop T. T
is determined by the extrapolation procedure for every time step in
the solidication interval. The heat transfer coefcient too can then
be calculated directly through Eqs. ((2)(4)) for every time step.
All assumptions that have been made for the identication of
the heat transfer coefcient are summarized below:

(3)

The ux q s is assumed to decrease linearly in the solidication


interval:
q s =

2(1 ((t tliquid )/(tsolid tliquid )))s x Es


t

(tliquid t tsolid )(4)

where s is the density of the melt, x is half the thickness of the


plate, cp is the heat capacity of the melt, T(t) is the temperature
at the time t, t is the time increment, Es is the latent heat of the
alloy, tliquid is the beginning of solidication and tsolid is the end of
solidication. The material parameters for IN738LC entered in the
calculations are given in Table 3.
Table 3
Material parameters for IN738LC taken from literature (Chapman et al., 2008).
s
cp
Es
t

7400
800
256
0.5

kg/m3
J/(kg K)
kJ/kg
s

Fig. 3. Determination of the temperature drop T at the interface alloy/mold. Example for time step t = 20 s.

184

C.H. Konrad et al. / Journal of Materials Processing Technology 211 (2011) 181186

Fig. 4. Specic heat capacity of the ceramic mold.

Heat is only transported perpendicular to the mold surface.


The heat transfer in the melt is innitely high.
The temperature propagation in the mold is linear because of constant heat conductivity in the mold. Therefore the temperature
of the mold at the interface can be extrapolated.
The solidication energy decreases linearly during freezing.
4. Results
4.1. Specic heat capacity
With the Linseis L81/1550 DSC a mean heat capacity of
1.15 J/(g K) between 200 and 1200 C was measured for the ceramic
of the mold. With the Netzsch 404 Pegasus DSC a slightly lower
heat capacity of 1.12 J/(g K) was measured. Both measurements
coincide well with values presented by Chapman et al. (2008) for
ceramic molds with similar composition (ZrO2 Al2 O3 SiO2 ). Fig. 4
shows the specic heat capacity of ceramic mold versus temperature.
4.2. Thermal diffusivity
Fig. 5 plots the thermal diffusivity over temperature in comparison with literature data. An estimated error bar is shown
exemplarily for a measurement at 600 C. The thermal diffusivities are in very good agreement with the results of Chapman et al.
(2008) for a similar ceramic mold material.

Fig. 6. Thermal conductivity of ceramic mold.

4.4. Temperature propagation


Fig. 7 shows a temperature prole with the extrapolated values at the alloy/mold interface. At time t = 0 the alloy is poured
into the mold. Negative times are plotted to show the cooling of
the mold while it is lifted from the heating to the pouring position. In these 4 s the mold cools down from 1000 to about 900 C at
the outside, but remains at 1000 C at the inner surface. When the
mold is completely lled the central thermocouple at 1 mm gives
a temperature of the still liquid alloy of 1410 C. The temperature
in the mold-wall increases rapidly within the rst seconds, due to
the heat transfer from the melt. The maximum temperature at the
mold/alloy interface of 1247 C is reached after 26 s. Thereafter the
temperature of the mold decreases steadily.
Solidication is apparently completed after about t = 60 s when
the cooling curves of the alloy kinks downward (see Fig. 7). The
remaining temperature drop at the interface between alloy and
mold is most probably due to a microscopic gap between mold
and solidied surface. The extrapolated temperature drop remains
constant until a macroscopic gap opens between alloy and mold
caused by different thermal expansions.
4.5. Liquidus and solidus temperature
Liquidus temperatures were identied at the intersection points
of two tangents to the concave kinks of the cooling curves (see
Fig. 7). Solidus temperatures were identied at the intersection
points of two tangents to the rst convex kinks of the cooling
curves. The average liquidus and solidus temperature from the
experiments are given in Table 4.

4.3. Thermal conductivity


Thermal conductivity  values calculated according to Eq. (1)
are presented in Fig. 6. Errors in thermal conductivity, density and
specic heat multiply. The resulting total error is again indicated
for the value at 600 C.

4.6. Grain size and secondary dendrite arm spacing


Figs. 8 and 9 show the expected behavior that grain size and
secondary dendrite arm spacing increase with mold pre-heating
temperature and mold thickness due to slower cooling rates.
4.7. Heat transfer coefcient
The temperature dependent heat transfer coefcient is shown
in Fig. 10 for the test T1000L6. For all experiments the heat transfer
coefcient is higher at the upper end of the solidication interval
when the alloy is mostly liquid and becomes lower approaching the
end of solidication. The mean heat transfer coefcient in the solidTable 4
Liquidus and solidus temperatures of IN738LC.

Fig. 5. Thermal diffusivity of ceramic mold in comparison with literature.

Tliq in C
Tsol in C

Experiment

Liu et al. (1998)

Chapman (2004)

1329 4
1278 9

1330
1282

1317
1233

C.H. Konrad et al. / Journal of Materials Processing Technology 211 (2011) 181186

185

Fig. 7. Temperature prole during casting (T1000L4).

5. Discussion

Fig. 8. Grain size and secondary dendrite arm spacing as functions of the mold
pre-heating temperature.

The specic heat capacity and the thermal diffusivity of the


ceramic shell mold system measured in this study were 1.1 J/(g K)
and 4.6 107 m2 /s, respectively. These values are in the same
range as the results given by Chapman et al. (2008) for similar shell
mold systems.
Liquidus and solidus temperature are in very good agreement
with DTA measurements from Liu et al. (1998) (see Table 4). On the
other hand Chapman (2004) published 12 and 45 C lower liquidus
and solidus temperatures, respectively. However, Liu et al. (1998)
pointed out that variations of the boron content typically encountered in commercial grade alloys may have a great inuence on the
liquidus and solidus temperature.
The measured grain sizes and secondary dendrite arm spacings
show the expected behavior. Higher mold temperatures and thicker
molds result in retarded solidication and thus in larger grains and
larger secondary dendrite arm spacing. The absolute values for the
grain sizes are generally four times lower than expected from the

Table 5
Heat transfer coefcients (outlier in italics).
Experiment

HTC in W/(m2 K)
Mean

Low

High

T1000L4
T1000L6
T1000L8
T900L6
T1100L6

560
430
500
250
470

310
300
410
180
410

660
500
530
300
550

Average (without T900L6)

490

360

560

ication interval as well as the values at the beginning and the end
of the solidication interval are listed in Table 5 for all experiments.
The low heat transfer coefcient for T900L6 of 250 W/(m2 K) is considered as an outlier, even though no experimental error could be
identied.

Fig. 9. Grain size and secondary dendrite arm spacing as functions of the mold
thickness.

186

C.H. Konrad et al. / Journal of Materials Processing Technology 211 (2011) 181186

dendrite arm spacing is simulated. Their results, which are in good


agreement with the experimental observations, are submitted for
publication in Metallurgical and Materials Transactions A.
6. Summary

Fig. 10. Heat transfer coefcient for 1000 C preheating temperature and 6 layers
(T1000L6).

The presented fairly simple direct method allows determining


the heat transfer coefcient in investment castings with high precision. The large scatter of mold material properties of investment
casting molds do not affect the result of the heat transfer coefcient
obtained by this method, because these properties do not enter the
calculations. Additionally, the experimental setup gives access to
all kind of temperatures that occur in the melt and mold and hence
allow a very good control of the casting process.
References

results of Liu et al. (1998) for a melt temperature of 1500 C without homogenization. Hence the grain sizes observed in this study
clearly reect the grain rening capability of the primary mold layer
containing CoAl2 O4 .
The heat transfer coefcients between 430 and 560 W/(m2 K) are
within the range of values published in literature for nickel based
superalloy melts in contact to ceramic shell molds. For example
Sahai and Overfelt (1995) give values of around 300 W/(m2 K) for
cylindrical geometries and of 505000 W/(m2 K) for plate geometries of the nickel based superalloy IN718.
All experiments show higher heat transfer coefcient values
at the beginning of solidication than at the end. A declining
heat transfer coefcient during solidication was also observed by
Browne and OMahoney (2001) as well as by Sahai and Overfelt
(1995).
Improvements of the presented direct method can be expected
by placing the thermocouples directly between the individual
shells of the mold during the manufacturing process. This would
allow placing the thermocouples closer to the interface and aligning
them more parallel to the isotherms in the mold. Also the errorprone drilling of the mold could be avoided.
The material parameters determined in this work are used by
Franke et al. (submitted for publication) for case studies of the
investment casting process with the commercial software ProCast.
There the correlation of solidication conditions with secondary

Aweda, J.O., Adeyemi, M.B., 2009. Experimental determination of heat transfer coefcients during squeeze casting of aluminum. Journal of Materials Processing
Technology 209, 14771483.
Browne, D.J., OMahoney, D., 2001. Interface heat transfer in investment casting of
aluminum alloys. Metallurgical and Materials Transactions A 32A, 30553062.
Chapman, L.A., 2004. Application of high temperature DSC technique to nickel based
superalloys. Journal of Materials Science 39, 72297236.
Chapman, L.A., Morrell, R., Quested, P.N., Brooks, R.F., Chen, L.-H., Ford, D., 2008.
PAMRIC: Properties of Alloys and Molds Relevant to Investment Casting, NPL
Report MAT 9. National Physics Laboratory, UK.
Franke, M.M., Hilbinger, R.M., Konrad, C.H., Glatzel, U., Singer, R., submitted for publication. Numerical determination of secondary dendrite arm spacing for IN738LC
investment castings. Metallurgical and Materials Transactions A 02.07.2010.
Krishnan, M., Sharma, D.G.R., 1996. Determination of the interfacial heat transfer
coefcient h in unidirectional heat ow by Becks nonlinear estimation procedure. International Communications in Heat and Mass Transfer 2, 203214.
Liu, L., Zhang, R., Wang, L., Pang, S., Zhen, B., 1998. A new method of ne grained
casting for nickel-base superalloys. Journal of Materials Processing Technology
77, 300304.
OMahoney, D., Browne, D.J., 2002. Use of experiment and an inverse method to study
interface heat transfer during solidication in the investment casting process.
Experimental Thermal and Fluid Science 22, 111122.
Ozisik, M.N., Orlande, H., Hector Jr., L.G., Anyalebechi, P.N., 1993. Inverse problem
of estimating interface conductance during solidication via conjugate gradient method. In: Gceri, S. (Ed.), First International Conference on Transport
Phenomena In Processing. Technomic Publishing Company, Inc, pp. 250265.
Reed, R.C., 2006. Superalloys. Cambridge University Press.
Sahai, V., Overfelt, R.A., 1995. Contact conductance simulation for alloy 718 investment castings of various geometries. Transaction of the American Foundrymens
Society 103, 627632.
Weathers, J., Johnson, A., Luck, R., Walters, K., Berry, J.T., 2006. Transiente Temperaturmessung. Giesserei-Praxis 10.

Das könnte Ihnen auch gefallen