Você está na página 1de 7

S. Y.

Kim1
e-mail: seoykim@kist.re.kr Assoc. Mem. ASME

Flow and Heat Transfer Correlations for Porous Fin in a Plate-Fin Heat Exchanger
The present experimental study investigates the impact of porous ns on the pressure drop and heat transfer characteristics in plate-n heat exchangers. Systematic experiments have been carried out in a simplied model of a plate-porous n heat exchanger at a controlled test environment. The porous ns are made of 6101 aluminum-alloy foam materials with different permeabilities and porosities. Comparison of performance between the porous ns and the conventional louvered ns has been made. The experimental results indicate that friction and heat transfer rate are signicantly affected by permeability as well as porosity of the porous n. The porous ns used in the present study show similar thermal performance to the conventional louvered n. However, the louvered n shows a little better performance in terms of pressure drop. For compactness of the heat exchanger, the porous ns with high pore density and low porosity are preferable. Useful correlations for the friction factor and the modied j-factor are also given for the design of a plate-porous n heat exchanger. S0022-14810001103-8 Keywords: Experimental, Finned Surfaces, Heat Exchangers, Heat Transfer, Porous Media

J. W. Paek B. H. Kang
Thermal/Flow Control Research Center, Korea Institute of Science and Technology, Seoul 130-650, Korea

Introduction
Heat exchange processes are indispensable to modern thermal engineering applications. On this account, tremendous studies on heat exchangers have been performed to enhance thermal transport between uids during past several decades 1. Recent studies have concentrated on the development of high-performance compact heat exchangers to reduce energy consumption as well as material cost. Various types of compact heat exchangers have been developed for specic applications. To exchange thermal energy between gas and liquid especially, compact heat exchangers with extended surfaces in the gas-side, such as wavy n, offset strip n, and louvered n, are often utilized. Such ns substantially reduce the gas-side thermal resistance, which always acts as the major factor of degrading the performance of plate-n heat exchangers. To date, the louvered n is known to be the most efcient and effective surface geometry in view of enhanced heat transfer rate and it is being widely used in automotive and aircraft air-cooled heat exchangers 2,3. However, the manufacturing process of a louvered n is rather complicated in that it requests a heavy nancial investment for maintenance. Furthermore, the louvered n has another potential disadvantage in structural strength since it is made of a very thin aluminum sheet. Consequently, there has been a demand for a new compact heat exchanger with ns that have a high heat transfer rate, and structural strength as well as a simple manufacturing process. In an attempt to respond to the demand, we have focused on the utilization of a porous metal n in place of the conventional louvered n. The impetus is a high-surface-area-to-volume ratio as well as enhanced ow mixing due to the tortuous path of the porous n. In addition, it has excellent characteristics in the structural strength as well as the simple manufacturing process of
1 To whom correspondence should be addressed. Thermal/Flow Control Research Center, Korea Institute of Science and Technology, P.O. Box 131, Cheongryang, Seoul 130-650, Korea. Contributed by the Heat Transfer Division for publication in the JOURNAL OF HEAT TRANSFER. Manuscript received by the Heat Transfer Division, June 24, 1999; revision received, March 10, 2000. Associate Technical Editor: M. Hunt.

metal foaming 4. Therefore, it is expected that the overall performance of plate-n heat exchangers can be improved by using the porous n. The literature surveyed reveals that much research on the porous materials can be found, mainly in energy-related problems, which include the utilization of geothermal energy, the control of pollutant spread in ground water, high performance insulation, etc. 5. Recently several papers have emphasized substantial heat transfer enhancement by using porous inserts in a ow passage 611. However, efforts to get ow and heat transfer data of the porous n for the design of compact heat exchanger are not being accomplished. In the present study, a simplied model of a plate-n heat exchanger, as shown in Fig. 1, is selected to investigate the impact of the porous n on a plate-n heat exchanger in detail. The simplied model emulates one ow passage of a conventional plate-n heat exchanger. A porous n is placed inside a channel, in which two channel walls are maintained at a constant temperature. The correlations of friction factor f and the j-factor for the porous n will be sought to provide a design parameter for heat exchangers. Further, we scrutinize the performance of the porous n in a plate-n heat exchanger, compared with that of the louvered n.

Fig. 1 Schematic conguration of the problem

572 Vol. 122, AUGUST 2000

Copyright 2000 by ASME

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 03/04/2014 Terms of Use: http://asme.org/terms

Fig. 2 Experimental apparatus

wall surface temperature. At steady state, the temperature deviation along the hot walls was less than 0.2C. All the channel walls were thermally insulated. To measure the inlet air temperature, three copper-constantan thermocouples Omega AWG36 were used at 5 mm upstream of the test section. Five copper-constantan thermocouples Omega AWG36 were also horizontally distributed at 5 mm downstream of the test section to measure the air temperature at the outlet. The downstream air temperatures measured at the core region in the channel were slightly lower than those at the wall region. For a high ow rate, the maximum deviation between the thermocouples was about 6.2 percent. By averaging the temperatures measured upstream and downstream of the test section, the overall temperature rise of air was evaluated. The overall temperature rise of air was varied from 20.4C to 33.8C according to the ow rate and the specimen. Experiments were started by inducing air ow to the channel and by maintaining the water jackets at constant temperature after installing the test specimen to the channel. The temperatures were monitored during the heat-up time by a data acquisition system Yokogawa DR230. After reaching a thermal steady state, the temperature data were recorded on the storage device for data analysis.

Experimental Apparatus and Test Procedure


A schematic diagram of the experimental apparatus is shown in Fig. 2. The experiments were conducted in a channel fabricated of Plexiglas of height H 9.0 mm and width W 90.0 mm. Test specimens of aluminum-alloy porous ns were made of height H 9.0 mm, W 90.0 mm, and length L 30.0 mm as depicted in Fig. 2. Pressure taps were installed at 5 mm upstream and downstream of the test section. The outlet of the channel was open to the atmosphere while the inlet was connected to a calming chamber in which the compressed air at 4 atm was induced through a rotameter. The rotameter was carefully calibrated by a bubble ow meter. Frontal air velocities U i tested in the present experiments ranged from 0.48 m/s to 3.64 m/s, and the corresponding Reynolds number based on the n height H was ReH 270 2050. To measure the pressure drop through the test specimen, the precision manometer Dywer Microtector 1430 with a resolution of 1/100 mm in water height was used. Two hot-water jackets made of copper were mounted on both side walls of the channel to provide a constant wall temperature condition. The water circulation loop supplied hot water to the jackets, as seen in Fig. 2. The temperature difference between the wall surface and the inlet air was set at 35C during the experiments. The inlet air temperature was about 20C. Three clamps were used to reduce thermal contact resistance between the jacket wall and the test specimen. As the compression loading of clamps was increased, the outlet air temperature approached an asymptotic value that was considered as a condition for minimal contact resistance. Three copper-constantan thermocouples Omega AWG36 were also inserted between the jacket wall and the surface of the test specimen to verify the uniformity of the

Test Specimen and Data Reduction


The test specimens used in the present study are porous ns made of aluminum-alloy 6101 foam. The porous ns have three distinct pore densities, i.e., 10, 20, and 40 PPI pores per inch at the same porosity of 0.92, and four different porosities, i.e., 0.89, 0.92, 0.94, and 0.96 at the same pore density of 20 PPI, as described in Table 1. The surface-area-to-volume ratio of the porous n increases as the pore density PPI increases or the porosity decreases. The permeability K of the porous n gradually decreases with an increase in the pore density due to the increased bulk friction. It is noted that permeability of the porous n with 20 PPI is proportional to (1 ) 2 , which shows a peak at 0.94. It may be attributed to the change of microscopic pore shape according to the porosity of the porous ns 12. The permeabilities of the porous ns were determined by using the Forchheimerextended Darcy model after measuring the pressure drop through the porous n. The effective thermal conductivity of the present porous n is little affected by the change of pore density while it is linearly increased by the decrease of porosity 13. The friction factor f and the modied j-factor j * are dened to assess the pressure drop and the heat transfer rate of the ns: f j * Ac 2 P , Ao G2 c (1)

s h 2/3 Pr s j . G cC p

(2)

In Eq. 1, the entrance and the exit loss coefcients are neglected 14. A c indicates the minimum free ow area, A c A , where A

Table 1 Test specimens of porous ns Surface-area to-Volume Ratio Ao /V m2/m3 790 1720 2740 2020 1510 1240 Effective Thermal Conductivity k e W/mK 13 5.33 5.56 6.01 6.77 4.27 2.82

Porous Fin 1 2 3 4 5 6

Pore Density (PPI) 10 20 40 20 20 20

Porosity 0.92 0.92 0.92 0.89 0.94 0.96

Permeability K m2 2.36 10 7 1.07 10 7 7.15 10 8 8.96 10 8 1.30 10 7 1.16 10 7

Darcy Number Da K / H 2 2.85 10 3 1.30 10 3 8.63 10 4 1.08 10 3 1.57 10 3 1.41 10 3

Journal of Heat Transfer

AUGUST 2000, Vol. 122 573

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 03/04/2014 Terms of Use: http://asme.org/terms

is the frontal area of the porous n, A WH . A o is the total heat transfer area of porous ns, i.e., A o A f A b , where A f and A b are the n area and the base area, respectively. is the density and P the pressure drop through the ns. G c is the mass velocity, G c U i , where U i is the frontal velocity. In Eq. 2, the modied j-factor includes the surface efciency s of the n to compare the thermal performance of the porous n with that of the conventional louvered n. Since it is difcult to evaluate the surface efciency of the porous n, we have multiplied the known value of the surface efciency for the conventional louvered n to its j-factor correlation 15. The surface efciency s of the louvered n is obtained from

sure drop and the mass velocity were, respectively, 2.8 percent and 1.5 percent at 95 percent condence. Then, the uncertainty in the friction factor was about 4.6 percent. The uncertainty of the modied j-factor that can be derived from Eq. 7 is

j* j*


T T
2

Tm Tm

Ao Ao

2 Pr 2 . (11) 3 Pr

s 1

Af 1 f . Ao

(3)

At Re1530, for example, the inlet, the outlet, and the wall temperatures were varied within 0.08, 0.16, and 0.22C at 95 percent condence, respectively. Therefore, the uncertainties in T and T m were, respectively, 0.6 percent and 2.0 percent in 95 percent condence level. The uncertainty in the Prandtl number ( Pr/Pr) was 0.7 percent. Consequently, the uncertainty in the modied j-factor was estimated by 2.4 percent.

The n efciency f is calculated from 16 tanh ml , f ml where m (4)

Results and Discussion


Before presenting the results for the porous ns, it is necessary to conrm the appropriateness of the experimental test model. Thus, the benchmark test by using a conventional louvered n has been performed, and the friction and j-factor values experimentally obtained were compared with the previous friction and j-factor correlations for the louvered ns in Fig. 3 18,3,15. The louvered n used in the present benchmark test has a dimension of louver pitch L p 1.0 mm, louver length L i 8.0 mm, n height H 9.0 mm, n pitch F p 1.88 mm, ow depth F d 26 mm, and n thickness t 0.1 mm. Surface-area-to-volume ratio of the louvered n is 1320 m2/m3. In Fig. 3, the friction factors of the present louvered-n model are estimated to be comparable to the Achaichia and Cowells correlation 18 for ReLp200. The correlation presented by Chang et al. 3 yields slightly lower values than the present louvered n data. It may be attributed that the n height of the louvered ns used in Chang et al. 3 is about two times larger than the present louvered n. However, the trend of the friction factors according to the Reynolds number based on the louver pitch is very similar. The j-factors of the present louvered n show a good agreement with the correlation by Chang and Wang 15. Consequently, it is believed that the present experimental model for ns is satisfactory to estimate the n performance of a plate-n heat exchanger with reasonable accuracy. Figure 4 shows the friction factor f for the porous ns as a function of the Reynolds number. The Reynolds number is based on the n height H and the porosity is 0.92. For comparison, the friction factor of the louvered n tested in the present study is also

2h . kt

(5)

Here, the n length l is the half of the channel height dened by l H /2 t . The t is the n thickness and k the thermal conductivity of n. The calculated surface efciency of the louvered n used in the present study was in the range of 0.930.98. The h denotes the convective heat transfer coefcient for the air side that is determined from the overall energy balance without considering contact thermal resistance:

sh U

G cC pA T . A o T m

(6)

In the present model of the porous n, therefore, s h equals the overall heat transfer coefcient U. Then, the modied j-factor from Eq. 2 is expressed as j * A T 2/3 Pr . Ao Tm (7)

In Eq. 7, the temperature difference of air between the inlet and the outlet T and the logarithmic mean temperature difference LMTD T m are T T o T i , T m T , ln T w T i / T w T o (8) (9)

where T i , T o , and T w denote, respectively, the air inlet, the air outlet, and the wall temperatures. The uncertainties in the present experimental results were estimated by the single-sample experiment analysis described by Kline and McClintock 17. The uncertainties for the friction factor and the modied j-factor were mainly attributed to the variation of temperature, velocity, pressure, and the thermophysical properties of uid while the effect of geometric uncertainty of the test section was meager. Therefore, the uncertainty of the friction factor can be expressed by

f
f

P P

Ao Ao

Gc
Gc

. (10)

In Eq. 10, the uncertainties of the porosity and the total surface area A o were 1.1 percent. The density variation was generally less than 1.5 percent because the friction factor was evaluated from the results of the cold-ow experiment. The uncertainties in the pres574 Vol. 122, AUGUST 2000

Fig. 3 Comparison of the friction and j -factors for the present louvered n with the previous correlations

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 03/04/2014 Terms of Use: http://asme.org/terms

Fig. 6 Friction factor correlation of the porous ns Fig. 4 Effect of pore density on the friction factors of the porous n at 0.92

plotted in the gure. The friction factor f decreases gradually as Re increases. The friction factor f of the high permeable porous n of 10 PPI is much higher than that of the lower permeable porous ns. This is attributed to the relatively small surface area of the high permeable porous n. It is also noted that the friction factor of the louvered n is slightly higher than that of the porous ns at low Reynolds numbers. As ReH increases, however, the porous ns are shown to have higher friction factors compared to the louvered n. The effect of porosity variation of the porous n on the friction factor is also displayed in Fig. 5. The pore density is xed at 20 PPI. The porous n of high porosity 0.96 shows a higher friction factor. Although the dimensional pressure drop decreases with the increase of the porosity 0.92 11, the friction factor dened in Eq. 1 increases due to the relatively rapid decrease in the surface area. It is also noted that the porous ns of porosity 0.96 have similar friction factors without displaying any monotonic decrease. It is caused by nonlinear behavior of the pressure drop permeability according to the variation of porosity, as mentioned in Table 1.

To derive a correlation of friction factors for the porous ns, the f data in Figs. 4 and 5 are converted by using the nondimensional parameters such as Da1/2, A c / A o and L / H , as demonstrated in Fig. 6. Then, it is found that the friction factors f for the porous ns tested in the present study generally merge to a single curve with maximum 15 percent deviation: f 2 Ac L 0.21 A c L . ReH Da A o H Da1/2 A o H (12)

Here, Da denotes the Darcy number K / H 2 . If the porous n is considered as a local volume averaged continuous medium, the above denition of the friction factor can be expressed in terms of the pressure drop in a smooth porous channel without ns. Rearranging the above equation yields
P/L H

U i2

0.105 1 . Re Da Da1/2

(13)

Fig. 5 Effect of porosity on the friction factors of the porous n at 20 PPI

It is interesting to note that Eq. 13 is exactly the same as the Forchheimer-extended Darcy model for porous media with an inertia coefcient of C E 0.105 5. In Eq. 13, the inertia coefcients C E of the present porous ns are slightly varied from 0.095 to 0.115 according to the pore density and the porosity of the n. Now we shall deal with the modied j-factor to show the heat transfer characteristics of the porous ns. Figure 7 displays the effect of pore density on the modied j-factor j * of the porous ns as a function of Reynolds number. The porosity is 0.92. As ReH increases, the modied j-factors are decreased and the corresponding convective heat transfer rates h are increased. This is in accordance with the previous results by Huang and Vafai 7, Amiri and Vafai 19, and Kim et al. 11. The modied j-factors are substantially pronounced as the pore density of the porous n becomes small. It is also observed that the modied j-factor of the louvered n used in the present study shows a similar value to that of the porous n with 10 PPI. The effect of porosity of the porous n on the modied j-factor j * is also exhibited in Fig. 8. The pore density is xed at 20 PPI. The modied j-factor values of the porous n with 0.96 are higher than those with lower porosity. For the ns of low porosity 0.94, the modied j-factors show similar values. This may be caused by the combined effect of the surface area and the effective thermal conductivity according to the change of the porosity see Table 1. In an attempt to get a modied j-factor correlation of the porous ns, the nondimensional parameter, the Darcy number Da, is emAUGUST 2000, Vol. 122 575

Journal of Heat Transfer

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 03/04/2014 Terms of Use: http://asme.org/terms

Fig. 7 Effect of pore density on the modied j -factors of the porous ns at 0.92

Fig. 9 Modied j -factor correlation of the porous ns

ployed again. When the modied j-factor data in Figs. 7 and 8 are converted using this nondimensional parameter, the data merge to a single line in the log-log plot exhibited in Fig. 9. Then, it gives a simple modied j-factor correlation with maximum 17 percent deviation for 270 ReH2050:
0.489 j * 13.73 ReH Da0.451 .

(14)

The above equation implies that thermal performance of Al-alloy porous ns can be easily predicted from measuring permeability K at a known ow rate. It should be noted here that the above modied j-factor correlation has similar functional relationship to that for conventional heat exchangers 3: j * C Rem n , (15) where C denotes a constant and is the nning factor that reects the ow geometry effects of the n. Thus, it is obvious that the Darcy number Da in Eq. 14 implicates the equivalent nning factor of the porous n in heat exchangers. Compactness of a heat exchanger is another important design parameter. Therefore, the volume goodness factor will be considered here. The air-side performance of the porous n for a unit

volume is expressed by s h . Here we should distinguish the air-side performance from the aforementioned modied j-factor. While the modied j-factor shows the heat transfer from a unit heat transfer surface, the air-side performance represents the total heat transfer that can be obtained in a unit heat-exchanging volume. Therefore, the porous n with low j-factors can show a high air-side performance when it has a high-surface-area-to-volume ratio. Also, the friction power consumption per a unit volume can be evaluated by 14 E U iA c P A o . Ao V (16)

Figure 10 delineates the air-side performance s h and friction power E characteristics for the porous ns at 0.92. The airside performance s h of the porous n increases with an increase in the friction power E . As the pore density PPI increases, the s h at the same friction power is pronounced. At a low friction power range, the s h of the louvered n is higher than that of the porous ns. When the friction power exceeds E 103 , however, the s h of the porous n with 40 PPI is

Fig. 8 Effect of porosity on the modied j -factors of the porous ns at 20 PPI

Fig. 10 Effect of pore density on the air-side performance s h of the porous ns at 0.92

576 Vol. 122, AUGUST 2000

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 03/04/2014 Terms of Use: http://asme.org/terms

Nomenclature
A frontal area of test specimen m2, A WH A b heat transfer area at the base of porous n m2, A b WL A c minimum free-ow area m2, A c A A f n surface area of porous n m2 A o total heat transfer surface area of porous n m2, A o A fAb C p specic heat at constant pressure J/kgK Da Darcy number, K / H 2 D h hydraulic diameter m, 2 H E friction power, Eq. 16 f friction factor, Eq. 1 h length-averaged heat transfer coefcient, Eq. 3 W/m2K H height of channel m j Colburn j-factor, j ( h / G c C p )Pr2/3 j * modied j-factor, Eq. 2 k thermal conductivity of n K permeability of porous n m2 L length of porous n m L p louver pitch of louvered n m PPI pores per inch Pr Prandtl number, / ReH Reynolds number based on the n height H , U i H / ReLp Reynolds number based on the louver pitch of louvered n, U i L p / T i inlet air temperature K T o outlet air temperature K T w wall temperature K U i frontal air velocity m/s W width of channel m Greek Symbols

Fig. 11 Effect of porosity on the air-side performance s h of the porous ns at 20 PPI

comparable to that of the louvered n. Therefore, it may be more benecial to use the low permeable porous n at the high-volume ow range. Inuence of porosity of the porous n on the air-side performance s h and the friction power consumption E is also depicted in Fig. 11. As the porosity decreases for a xed pore density of 20 PPI, the s h of the porous n increases. Therefore, it may be concluded that the porous ns with low permeability and low porosity are preferable in terms of the compactness of heat exchangers.

Conclusion
An experimental investigation on the effect of porous n in a simplied model of plate-n heat exchanger has been performed. Six aluminum-alloy porous ns of various permeabilities and porosities were selected to scrutinize the performance of porous ns for the application to plate-n heat exchangers. Comparison of heat transfer and friction characteristics between the porous ns and the conventional louvered n was made. Experimental results indicate that the friction factor is much lower for low permeable porous ns due to relatively larger surface area. The louvered n exhibits slightly higher friction factor values than that of the porous ns at low Reynolds numbers. When the Reynolds number is high, however, the porous ns show much higher friction factors compared to the louvered n. The modied j-factors of the porous ns decrease as the pore density increases or as the porosity decreases. It is noted that the present porous ns have a similar thermal performance compared to the conventional louvered n; however, the louvered n shows a little better performance in terms of pressure drop. The friction and heat transfer correlations for the porous ns have been obtained by employing the Darcy number, Da, and the geometrical parameters such as A c / A o and L / H . It is also found that the porous ns with low permeability and low porosity are preferable to compactness of plate-porous n heat exchangers.

heat transfer surface area to volume ratio m2/m3, Ao /V P pressure drop of air ow Pa T temperature difference of air between the inlet and the outlet K, Eq. 4 T m log mean temperature difference K, Eq. 5 porosity density of air kg/m3 kinematic viscosity m2/s f n efciency s surface efciency

References
1 Webb, R. L., 1994, Principles of Enhanced Heat Transfer, John Wiley and Sons, New York. 2 Sahnoun, A., and Webb, R. L., 1992, Prediction of Heat Transfer and Friction for the Louver Fin Geometry, ASME J. Heat Transfer, 114, pp. 893 900. 3 Chang, Y.-J., Wang, C.-C., and Chang, W. J., 1994, Heat Transfer and Flow Characteristics of Automotive Brazed Aluminum Heat Exchangers, ASHRAE Trans., 100, No. 2, pp. 643652. 4 Gibson, L. J., and Ashby, M. F., 1997, Cellular Solids, Cambridge University Press, Cambridge, UK. 5 Kaviany, M., 1991, Principles of Heat Transfer in Porous Media, SpringerVerlag, New York. 6 Hunt, M. L., and Tien, C. L., 1988, Effects of Thermal Dispersion on Forced Convection in Fibrous Media, Int. J. Heat Mass Transf., 31, pp. 301309. 7 Huang, P. C., and Vafai, K., 1993, Flow and Heat Transfer Control Using a Porous Block Array Arrangement, Int. J. Heat Mass Transf., 36, pp. 4019 4032. 8 Huang, P. C., and Vafai, K., 1994, Analysis of Forced Convection Enhancement in a Channel Using Porous Blocks, J. Thermophys. Heat Transfer, 8, pp. 563573. 9 Hadim, A., 1994, Forced Convection in a Porous Channel With Localized Heat Sources, ASME J. Heat Transfer, 116, pp. 465472. 10 Sung, H. J., Kim, S. Y., and Hyun, J. M., 1995, Forced Convection From an Isolated Heat Source in a Channel With Porous Medium, Int. J. Heat Fluid Flow, 16, pp. 527535. 11 Kim, S. Y., Kim, J.-H., and Kang, B. H., 1998, Effect of Porous Fin in a

Acknowledgment
This work was supported by a grant from the Critical Technology 21 Project of the Ministry of Science and Technology, Korea and Jang-Han Engineering, Inc. The support of test specimen from ERG, Inc. is deeply acknowledged by the authors. Journal of Heat Transfer

AUGUST 2000, Vol. 122 577

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 03/04/2014 Terms of Use: http://asme.org/terms

Plate-Fin Heat Exchanger, Proceedings of the ASME Heat Transfer Division, 3, ASME, New York, pp. 477482. 12 ERG Duocel Aluminum Foam Catalog, 1995, Energy Research and Generation, Inc. Oakland, CA. 13 Paek, J. W., Kang, B. H., Kim, S. Y., and Hyun, J. M., 2000, Effective Thermal Conductivity and Permeability of Aluminum Foam Materials, Int. J. Thermophys., 21, pp. 453464. 14 Kays, W. M., and London, A. L., 1984, Compact Heat Exchangers, McGrawHill, New York. 15 Chang, Y.-J., and Wang, C.-C., 1997, A Generalized Heat Transfer Correlation for Louver Fin Geometry, Int. J. Heat Mass Transf., 40, pp. 533534.

16 McQuiston, F. C., and Parker, J. D., 1994, Heating, Ventilating, and AirConditioning Analysis and Design, John Wiley and Sons, New York. 17 Figliola, R. S., and Beasley, D. E., 1995, Theory and Design for Mechanical Measurements, Wiley, New York. 18 Achaichia, A., and Cowell, T. A., 1988, Heat Transfer and Pressure Drop Characteristics of Flat Tube and Louvered Plate Fin Surfaces, Exp. Therm. Fluid Sci., 1, pp. 147157. 19 Amiri, A., and Vafai, K., 1994, Analysis of Dispersion Effects and NonThermal Equilibrium Non-Darcian, Variable Porosity Incompressible Flow Through Porous Medium, Int. J. Heat Mass Transf., 37, pp. 939954.

578 Vol. 122, AUGUST 2000

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 03/04/2014 Terms of Use: http://asme.org/terms

Você também pode gostar