You are on page 1of 7

Applied Thermal Engineering 30 (2010) 2815e2821

Contents lists available at ScienceDirect

Applied Thermal Engineering


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

Robust methodology for determination of heat transfer coefficient distribution


in convection
Bowang Xiao a, *, Qigui Wang b, Gang Wang a, Richard D. Sisson Jr. c, Yiming Rong a
a
Manufacturing Engineering, Worcester Polytechnic Institute, 100 Institute Road, Worcester, MA 01609, USA
b
General Motor Company, Materials Technology, Global Powertrain Engineering, 823 Joslyn Avenue, Pontiac, MI 48340, USA
c
Materials Science and Engineering, Worcester Polytechnic Institute, 100 Institute Road, Worcester, MA 01609, USA

a r t i c l e i n f o a b s t r a c t

Article history: To predict the microstructures, residual stresses and distortions in the heat treated metal components, it
Received 12 April 2010 is important to accurately know the heat transfer coefficients (HTCs) between the hot work piece and
Accepted 13 August 2010 cooling media. In this paper, a new method is presented to accurately determine the node-based HTC
Available online 20 August 2010
distribution by coupling computational fluid dynamics (CFD) with optimal weight functions and scale
factors. With this new method, the predicted temperature profile of the work piece during quenching
Keywords:
(rapid cooling) is in excellent agreement with experimental measurements. This new method can be also
Heat transfer coefficient
applied to accurately predict convection heat transfer in thermal equipment such as heat exchangers and
Convection
Quenching
refrigerators, building thermal design and other heat transfer related situations.
Finite element analysis Ó 2010 Elsevier Ltd. All rights reserved.
CFD

1. Introduction vT
Q_ cond ¼ KA (1)
vx
To improve mechanical properties, many heat treatable metal
products are usually subject to heat treatment that involves where Q_ cond ¼ heat transfer rate via conduction, W; K ¼ thermal
a quenching process to produce supersaturated solid solution or conductivity of a material, W/m  C; A ¼ cross section area, m2;
martensitic phase transformation. A significant amount of residual vT/vx ¼ thermal gradient in the direction of the heat flow,  C/m
stress can be developed in the quenched parts [1,2]. The existence
of residual stresses can have a significant detrimental influence on
Q_ conv ¼ hc AðT  TN Þ (2)
the performance of a structural component. In many cases, the high where Q_ conv ¼ heat transfer rate via convection, W; hc ¼ convection
tensile residual stresses can also result in a severe distortion of the heat transfer coefficient, W/m2  C; A ¼ surface area, m2;

component, and they can even cause cracking during quenching or T ¼ component surface temperature, C; TN ¼ quenchant tempera-

subsequent manufacturing processes [3,4]. The magnitude of ture, C
residual stress and distortion produced in the parts during  
quenching significantly depends on the cooling rate and the extent Q_ rad ¼ 3sA T 4  TN
4
(3)
of non-uniformity of the temperature distribution in the parts. The
heat transfer during quenching involves conduction as expressed in where Q_ rad ¼ heat transfer rate via radiation, W; s ¼ universal
Equation (1), convection as expressed in Equation (2), and radiation Stefan Boltzman constant, 5.6704  108 W/m2 K4; 3 ¼ emissivity of
as expressed in Equation (3). Experimental and numerical simula- the body.
tion results have shown that the convection HTC between the hot To simulate the quenching process and predict microstructure,
component and the quenching media plays an important role in residual stresses and distortions in the as-quenched parts, accurate
resultant distortion, residual stress and hardness distribution of the HTC distribution data are needed. Accurate HTC distribution data
quenched object [5e8]. are also required for optimizing designing thermal equipment such
as heat exchanges and refrigerators and thermal controls in
buildings and electrical products [9e11].
There are many classical empirical equations reported in the
* Corresponding author. Tel.: þ1 508 831 6160; fax: þ1 508 831 6412. literature for calculating convection HTC data by using dimension-
E-mail address: bowangxiao@gmail.com (B. Xiao). less numbers (i.e. Reynolds number, Prandtl number and Nusselt

1359-4311/$ e see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.applthermaleng.2010.08.017
2816 B. Xiao et al. / Applied Thermal Engineering 30 (2010) 2815e2821

Fig. 1. Comparison of measured and calculated timeetemperature curves using a CFD code at two locations of an aluminum casting quenched in water [22].

number) [12]. However, their applications are limited since almost iterative modification method, divides the part surfaces into several
all of these dimensionless numbers are calibrated for specific zones and optimizes the zone-based HTC distribution data iteratively
experimental conditions which can be significantly different from by minimizing the errors between the predicted temperature
the actual production process. An experimental approach with distributions and the measured cooling curves at different locations
a small probe can be utilized to determine HTC. With the so-called of the part [14,20]. Although this method can provide accurate
lumped heat capacity method [13,14], the cooling curves of the prediction of the temperature distributions of the locations where
probe are acquired and used to inversely calculate HTC in terms of the thermocouples are instrumented, it cannot assure that the pre-
probe surface temperature [12,14e18]. This method assumes that dicted temperature profiles at other locations in the part are
the temperature distribution in the small probe is uniform during matching the actual temperature distributions during quenching.
cooling in order to calculate HTC inversely, which is also uniform Another limitation of the method is that temperature measurements
over the probe surface [13]. Another method, known as the direct are needed for the same part quenched at different conditions.
conduction method, assumes the temperature near a surface is With the recent advances in CFD, the code may be used to
a linear variation so that the HTC can be obtained based on the calculate the node-based HTC distribution [21]. However, no
equilibrium of heat flux at the surface by conduction and convection currently commercially available CFD code is designed to calculate
[13,19]. It is, however, difficult to obtain HTC distribution data with accurate HTC distribution data. The current CFD prediction of heat
this experimental method. transfer and temperature distribution of a quenched part is not
In recent years, some methods have been reported to obtain HTC accurate particularly in liquid quenching because the complicated
distributions for the entire heat transfer interface. One method called interaction and heat transfer phenomena between liquid and the

START

FILE: HTC distribution data


exported from CFD simulation

INITIAL
MODIFY
Scale factors

Thermal simula tion in


ΔT >= 5 oC
ABAQUS and
Iteration #
YES <=10
Read Temperature from NO
ABAQUS results

FILE:
Experimental Calculate END
temperature vs. temperature
time curves differences ΔT

Fig. 2. Flow chart showing the procedure to optimize the HTC distribution (NOTE: the scale factors can vary with work piece temperature or quenching time) [23].
B. Xiao et al. / Applied Thermal Engineering 30 (2010) 2815e2821 2817

hot part are not fully understood and represented in the state-of-
the-art fluid flow and heat transfer code. Fig. 1 shows examples of
the significant discrepancy observed in the thermal simulation
using a state-of-the-art fluid flow and heat transfer code in
comparison with experimental measurements in a water-quenched
aluminum part [22].
In this paper, a new method to accurately determine node-based
HTC distributions by coupling computational fluid dynamics (CFD)
with optimal weight functions and scale factors is presented.

2. Integrated CFD and iterative modification method

Fig. 2 summarizes the integrated CFD and iterative modification


method for accurate determination of HTC distribution during
quenching [23]. An initial set of node-based HTC distribution data
are first obtained from the CFD simulation based on the part
geometry, quenching setup, and quenching conditions including
part initial temperature, material properties, quenchant flow
Fig. 4. A schematic illustration of the aluminum alloy casting with 14 thermocouples
velocity, quenching gas flow direction relative to the part, quen-
at selected locations.
chant temperature and physical properties. The initial HTC values
for the entire surface of the part calculated from the CFD simulation
are then optimized by multiplying scale factors to minimize the
modify the optimized baseline HTC distribution data for different
errors between the predicted and the measured temperatures for
quenching conditions (i.e., variations of quenching conditions such
a given quenching condition. Equations (4) and (5) illustrate how
as flow velocity from the baseline) without performing complete
the scale factors are modified based on the temperature differences
heat transfer and optimization calculations, as shown in Fig. 3. For
between the simulation and measurement. In the optimization
instance, a heat treating company has acquired some time-
process, the scale factors are modified iteratively until either
etemperature curves for a specific part under current production
acceptable temperature differences such as 5  C are reached or
condition and has performed the optimization presented in Fig. 2 to
iteration number reaches the preset maximal iteration number (to
obtain the HTC distribution data. Now the company increases the
prevent endless loop). It should be noted that the scale factors can
air velocity for a higher cooling rate. With the weight function for
vary with the surface temperatures of the part or quenching time. If
air velocity, the company can quickly determine the new HTC
the part surfaces are grouped into surface zones, different scale
distribution by scaling up the old one. In this way, no quenching
factors can be applied to various surface zones.
experiment is required for determining the new HTC distribution.
ScaleFactornew ¼ ScaleFactorold þ DScaleFactor (4) Neither is the CFD simulation for the new condition.

DScaleFactor ¼ k$DT (5)


where k is a constant and its value only affects the convergence
rate.
The calculation time in this optimization process varies with the
complexity of the target part surface, the initial guess and the value
of the constant k in Equation (5). For the optimization in the
following validation section, the HTC data converge usually after
about 10 iterations with a reasonable k. For a series of optimizations
with similar conditions, the optimized HTC data for the previous
condition can be used as an initial guess of the next optimization and
the optimization process can even be completed after only 1 or 2
iterations. The calculation time for each iteration depends on the
mesh used for the solid part, the time increment and the total cooling
time. For the aluminum casting shown in Fig. 4, it takes less than 1 h
to complete one iteration on a 4-CPU server by using Abaqus.
When the node-based HTC data are optimized for a baseline
quenching condition, a set of semi-empirical equations (or weight
functions) as expressed in Equation (6) can be used to quickly

CFD Scale factors HTC


simulation distribution

New New
Experiment HTC
production Semi-empirical
condition equations distribution

Fig. 3. A schematic illustration of the integrated CFD, iterative modification method Fig. 5. The experimental system for the forced air quenching of the aluminum alloy
and semi-empirical equations in determining node-based HTC distribution. casting.
2818 B. Xiao et al. / Applied Thermal Engineering 30 (2010) 2815e2821

modification factor and the air velocity for the velocity ranging from
4 m/s to 20 m/s, which was determined by experiments [24].
 
Vel
Kvelocity ¼ A  þB (7)
Vel0
where A ¼ 0.57; B ¼ 0.41; and Velo ¼ 10.5 m/s.

3. Validation and application of the method

To validate the method, a mixed leg thickness picture-frame


shape aluminum casting was utilized in the forced air quenching
experiments. Fig. 4 shows the casting instrumented with 14 ther-
mocouples. The dots and lengths of dash lines illustrate the locations
of the thermocouples in three dimensions. In air quenching exper-
iments, the casting was heated in a furnace to about 485  C and then
placed on the fixture and quenched in the forced air, as shown in
Fig. 5. The cooling curves at 14 locations of the casting during
Fig. 6. A sample timeetemperature profile at the vertical orientation and 18 m/s gas
quenching were acquired by a data acquisition system. Temperature
velocity.
measurements were taken 5 times per second. Fig. 6 shows one
exemplary timeetemperature profile for the aluminum casting
during quenching at the vertical orientation as shown in Figs. 7 and
HTC ¼ K1 K2 K3 /Kn $HTCo (6)
8. The air temperature, pressure and humidity were measured by
where HTCo is the standard HTC at a baseline quenching condition, a weather station (La Crosse Technology WS-7394U-IT-CH Wireless
unit in W/m2  C; K1, K2, K3, ., Kn are modification factors. Forecast Station). The air velocity at the quenching area was cali-
These modification factors are governing the effects of various brated as 18 m/s with an anemometer.
influencing factors such as fluid velocity, fluid flow direction, fluid A CFD simulation was conducted based on the experimental
temperature, part surface quality and material. These modification setup, as shown in Fig. 7. A big air box around the casting was
factors can be either constant or temperature-dependent. They can be created. Boundary conditions were applied to the air box. One face
determined experimentally or numerically by CFD simulations. For was defined as the air velocity inlet where normal air flow at
instance, Equation (7) shows the relationship between the velocity specified air velocity was blown in. All other faces were defined as

Fig. 7. CFD modeling.


B. Xiao et al. / Applied Thermal Engineering 30 (2010) 2815e2821 2819

Fig. 10. A comparison of temperatureetime curves between the measurements and


FEA simulation using the optimized HTC data (one simulation using original HTC for
comparison purpose).
Fig. 8. A color contour showing the CFD-predicted HTC distribution in the aluminum
test casting during forced air quenching [23].
varies from location to location even on the same surface of the
casting.
opening boundary where air can go in and out depending on the However, the predicted node-based HTC distribution directly
pressure difference. Keepsilon turbulence model and P1 radiation from the CFD simulation is not accurate. As shown in Fig. 9, the
model were applied in this CFD simulation. Simulation results simulated cooling curves from the finite element analysis using
showed that radiation slightly affects the temperature cooling the original CFD-determined HTC data significantly differ from the
curve. In this simulation, air is treated as ideal gas. experimentally measurements at the locations as shown in Fig. 4.
In the CFD simulation, stagnant air (initial status) was forced to Accordingly, the original CFD-determined HTC data are modified
flow and a steady air flow field was formed after a few seconds. In and optimized with the scale factor as described in Equations (4)
the simulation, fresh air was blown to the cooling zone and thus the and (5). After the HTCs are optimized with scale factors, the node-
inlet air temperature remained constant during simulation. Since based HTC distribution represents better heat transfer phenomena
the air flow (18 m/s, 1 atmospheric pressure) was set steady and air during quenching. As shown in Fig. 10, the predicted cooling curves
temperature remained constant, the calculated HTC distribution is using the optimized HTC data agree very well with the experi-
considered steady. Fig. 8 presents the predicted node-based HTC mentally measured temperature distributions. The maximum error
distribution from the CFD simulation when the air flow has become (or difference) between the predicted temperatures and the
steady. As expected, the predicted node-based HTC distribution measurements is þ/5  C. Fig. 10 also shows how the optimized
HTC data drag the timeetemperature curve at location TC2 towards
the experimental one and reduce the temperature differences from
above 50  C to less than 5  C. As presented in Fig. 11, the corre-
sponding scale factor varies with casting surface temperature and
the original HTC magnitude is scaled up by over 30%.
The optimization is further validated by comparing the cool-
ing rates. Fig. 12 compares the experimentally determined cool-
ing rates at TC2, TC3, TC5 and TC9 (on four different walls),
predicted cooling rate using CFD-determined HTC data at TC2,
and predicted cooling rate using optimized HTC data at TC2. The

Scale Factor
1.4
Scale Factor

1.3

1.2

1.1

1
0 100 200 300 400 500

Fig. 9. A comparison of temperatureetime curves between the measurements and FEA Temperature (oC )
simulation using the original CFD-determined HTC data (The differences between
experimental TC3 and TC5 are very small). Fig. 11. The scale factor as a function of part surface temperature.
2820 B. Xiao et al. / Applied Thermal Engineering 30 (2010) 2815e2821

Cooling Rate Comparison complicated. As shown in Fig. 9, the temperature differences


0 between the measurements and the predictions based on the HTC
obtained directly from CFD simulation can be as high as 50  C for
-1 the relatively simple shape aluminum test casting. Therefore, the
thermal calculation from CFD simulation cannot be directly used in
-2 structural analysis particularly for the residual stress and distortion
prediction.
2 measured
Cooling rate ( oC/s)

-3 To accurately simulate thermal history and predict residual


2 simulated (after stress and distortion in the quenched part, the node-based HTC
-4 optimization) distribution data determined from CFD simulation need to be
2 simulated (before optimized. The CFD-determined HTC distribution can be utilized as
optimization)
-5 an initial set of HTC data for further optimization. In the HTC
3 measured
optimization, the node-based HTC distribution data are modified
-6
5 measured iteratively with scale factors or weight functions (semi-empirical
equations). The scale factors and weight functions are not only
-7
9 measured temperature-dependent but also part surface geometry dependent.
With the measured cooling curves from the quenched part, the
-8
0 100 200 300 400 500 scale factors and/or weight functions can be calibrated and opti-
Surface Temperature ( oC ) mized for a given quenching condition. As shown in Fig. 10, the
temperature difference between the measurements and the
Fig. 12. A comparison of cooling rates among experimentally obtained and numeri-
cally predicted ones.
predictions based on the optimal HTC distribution can be less than
5  C.
Although this new method was initially put forward for
experimental cooling rates are generated from the experimental quenching processes, it can be also applied to other heat treating
cooling curves, which are moving averaged to remove abrupt processes like heating, annealing, tempering, etc., thermal design
oscillations. It is apparent that the predicted cooling rate using and management and others, whenever there is convection heat
optimized HTC data agrees very well with the experimental one, transfer and accurate prediction of temperature distribution in
but the predicted cooling rate using original CFD-determined HTC a part is required.
data differs from the experimental one significantly.
In addition, the cooling rate plot validates two other things. 5. Conclusion
First, at the beginning, because of the transportation of casting from
the furnace to the cooling zone, there are big changes of cooling In this paper, a new method integrating CFD simulation and an
rates at the beginning of cooling. Second, the cooling rates in the iterative modification and optimization method was developed to
main range are very linear, indicating that HTC data during gas determine accurate node-based HTC distribution for a part with
cooling remain quite constant after the gas flow becomes steady. complicated geometry during rapid cooling. With this method, the
This conclusion can be drawn from the following derivation. accurate HTC distribution data can be obtained and thus transient
In the quenching process, the heat loss rate of the hot part can heat transfer in the cooled part can be accurately predicted. When
be measured by Equation (8). further integrated with semi-empirical equations (weight func-
tions), this method can be used to determine new HTC distributions
Q_ loss ¼ Cp m$dT=dt (8) for the same part under different quenching conditions at no cost of
experiments and simulations.
where Q_ loss ¼ heat loss rate during cooling, W; Cp ¼ specific heat of
This method can be applied in other numerical predictions of
the casting, J/kg  C; m ¼ mass, kg; dT/dt ¼ cooling rate,  C/s.
convection heat transfer, such as optimization of heat exchanger
If we simplify the problem and assume the heat loss of a small
design, thermal management of electrical products, etc.
portion of the part is transferred by convection, since radiation is
very small in this case, we can approximate Equation (8) as Equation
Acknowledgements
(2). Thus, after substituting Equation (8) into Equation (2), we have
the cooling rate as a function of surface temperature as expressed in
This work was supported financially by GM Powertrain in
Equation (9). The near linear relationship between cooling rate and
Pontiac, Michigan. The authors thank GM R&D Foundry for
surface temperature implies that HTC is close to a constant.
producing the aluminum alloy castings. The gratitude is also sent to
hc A hc A CHTE (Center for Heat Treating Excellence) members for their
dT=dt ¼ T TN (9) sponsored quenching system.
Cp m Cp m
References

4. Discussion [1] I. Elkatatny, Y. Morsi, A.S. Blicblau, S. Das, E.D. Doyle, Numerical analysis and
experimental validation of high pressure gas quenching, Int. J. Therm. Sci.
42 (4) (2003) 417e423.
Although much commercial CFD software can solve thermal and [2] K. Li, B. Xiao, Q. Wang, Residual stresses in as-quenched aluminum castings,
fluid problems to a certain degree of accuracy, the software cannot SAE Int. J. Mater. Manuf. 1 (1) (2009) 725e731.
[3] P. Li, D.M. Maijer, T.C. Lindley, P.D. Lee, Simulating the residual stress in an
accurately calculate the heat transfer coefficients and temperature
A356 automotive wheel and its impact on fatigue life, Metall. Mater. Trans. B.
distributions during highly transient heat transfer processes like 38 (4) (2007) 505e515.
quenching. This is not only because many simplifications and [4] Y.L. Lee, J. Pan, R. Hathaway, M. Barkey, Fatigue Testing and Analysis: Theory
assumptions are assumed in the CFD calculations but also because and Practice. Elsevier Butterworth-Heinemann, 2005.
[5] A. Rose, O. Kessler, F. Hoffmann, H.W. Zoch, Quenching distortion of aluminum
the physics and the phase transformation phenomena involved at castings-improvement by gas cooling, Materialwiss. Werkst. 37 (1) (2006)
the interface between the hot part and cooling media are 116e121.
B. Xiao et al. / Applied Thermal Engineering 30 (2010) 2815e2821 2821

[6] Z. Li, R.V. Grandhi, R. Srinivasan, Distortion minimization during gas [16] D.K. Funatani, M. Narazaki, M. Tanaka, Comparisons of probe design and cooling
quenching process, J. Mater. Process. Technol. 172 (2) (2006) 249e257. curve analysis methods, in: 19th ASM Heat Treating Society Conference Proceed-
[7] H. Li, G. Zhao, S. Niu, Y. Luan, Technologic parameter optimization of gas ings Including Steel Heat Treating in the New Millennium, 1999, pp. 255e263.
quenching process using response surface method, Comput. Mater. Sci. 38 (4) [17] G.E. Totten, C.E. Bates, N.A. Clinton, Handbook of Quenchants and Quenching
(2007) 561e570. Technology. ASM International, Materials Park, OH, 1993.
[8] B. Xiao, G. Wang, Y. Rong, Hardenability and distortion control in high pres- [18] N.I. Kobasko, A.A. Moskalenko, G.E. Totten, G.M. Webster, Experimental
sure hydrogen quenching, Int. J. Manuf. Res., in press. determination of the first and second critical heat flux densities and quench
[9] B. Sahin, Y. Ust, I. Teke, H.H. Erdem, Performance analysis and optimization of process characterization, J. Mater. Eng. Perform. 6 (1) (1997) 93e101.
heat exchangers: a new thermoeconomic approach, Appl. Therm. Eng. [19] C. Laumen, S.I. Midea, T. Lubben, F. Hoffman, P. Mayr, Measured heat transfer
30 (2e3) (2010) 104e109. coefficients by using hydrogen as quenchant in comparison with helium and
[10] G.N. Xie, B. Sunden, Q.W. Wang, Optimization of compact heat exchangers by nitrogen, Accelerated Cooling/Direct Quenching of Steels, 1997, pp. 199e206.
a genetic algorithm, Appl. Therm. Eng. 28 (8e9) (2008) 895e906. [20] M. Li, J.E. Allison, Determination of thermal boundary conditions for the
[11] Z. Luo, H. Cho, X. Luo, K. Cho, System thermal analysis for mobile phone, Appl. casting and quenching process with the optimization tool OptCast, Metall.
Therm. Eng. 28 (14e15) (2008) 1889e1895. Mater. Trans. B 38B (4) (2007) 567e574.
[12] J.P. Holman, Heat Transfer, ninth ed. McGraw-Hill, New York, 2002. [21] G. Wang, B. Xiao, L. Zhang, K. Rong, Q. Wang, Heat transfer coefficient study
[13] M. Sedighi, C.A. McMahon, The influence of quenchant agitation on the heat for air quenching by experiment and CFD modeling, in: 137th Annual Meeting
transfer coefficient and residual stress development in the quenching of & Exhibition. Materials Characterization, Computation and Modeling, vol. 2,
steels, Proc. Inst. Mech. Eng. Pt. B: J. Eng. Manuf. 214 (7) (2000) 555e567. 2008, pp. 271e276.
[14] A. Sugianto, M. Narazaki, M. Kogawara, A. Shirayori, A comparative study on [22] O. Dahlberg, Z. Wu, Feasibility study of three dimensional quenching analysis,
determination method of heat transfer coefficient using inverse heat CD-adapco Technical Report to GM. 071-186-003, 2008.
transfer and iterative modification, J. Mater. Process. Technol. 209 (10) [23] Q. Wang, B. Xiao, G. Wang, Y. Rong, R.D.J. Sisson, Systems and methods for
(2008) 4627e4632. predicting heat transfer coefficients during quenching (Pending Patent), 2010.
[15] M. Maniruzzaman, J. Chaves, C. McGee, S. Ma, R. Sisson Jr., CHTE quench probe [24] B. Xiao, G. Wang, Q. Wang, M. Maniruzzaman, R.D. Sisson Jr., Y. Rong, An
system: a new quenchant characterization system, in: 5th International Confer- experimental study of heat transfer during forced air convection, J. Mater. Eng.
ence on Frontiers of Design and Manufacturing (ICFDM), 2002, pp. 619e625. Perform., in press, doi: 10.1007/s11665-010-9745-7.

You might also like