metal foam heat exchanger

Upload: muki10

Post on 08-Jan-2016

226 views

Category:

Documents


0 download

DESCRIPTION

Metal Foam

TRANSCRIPT

  • puli

    titute

    ETH Center, ML J 36, 8092 Zurich, Switzerlandb ABB Corporate Research Ltd., Segelhof, 5405 Baden-Daattwil, Switzerland

    pressure drop across a porous medium as a linear

    function of the ow velocity and material perme-

    ability. Since then, a series of improvements have

    (1918). The addition of a quadratic term in the

    linear Darcy Law was rst proposed by Dupuit

    (1863), although many mistakenly credit Forch-heimer with the addition of the quadratic term as a

    result of his extensive review of other porous me-

    dia studies (Forchheimer, 1901). A thorough re-

    Mechanics of Materials 35 (20

    ARTICLE IN PRESS* Corresponding author. Tel.: +41-1-632-2738; fax: +41-1-Keywords: Metal foam; Porous media; Heat transfer; Open-cell aluminum foam

    1. Introduction

    Flow through porous media has been studied in

    detail ever since Darcys publication in 1856(Darcy, 1856). His work described the uid ow

    been made to describe the pressure drop behavior

    in better detail, one of which was accounting for

    temperature variations in the uid by Hazen

    (1893). These temperature variations were later

    linked to the viscosity of the uid by KruugerReceived 1 June 2002; received in revised form 30 January 2003

    Abstract

    Open-cell metal foams with an average cell diameter of 2.3 mm were manufactured from 6101-T6 aluminum alloy

    and were compressed and fashioned into compact heat exchangers measuring 40.0 mm 40.0 mm 2.0 mm high,possessing a surface area to volume ratio on the order of 10,000 m2/m3. They were placed into a forced convection

    arrangement using water as the coolant. Heat uxes measured from the heater-foam interface ranged up to 688 kWm2,which corresponded to Nusselt numbers up to 134 when calculated based on the heater-foam interface area of 1600

    mm2 and a Darcian coolant ow velocity of approximately 1.4 m/s. These experiments performed with water were

    scaled to estimate the heat exchangers performance when used with a 50% waterethylene glycol solution, and werethen compared to the performance of commercially available heat exchangers which were designed for the same heat

    transfer application. The heat exchangers were compared on the basis of required pumping power versus thermal re-

    sistance. The compressed open-cell aluminum foam heat exchangers generated thermal resistances that were two to

    three times lower than the best commercially available heat exchanger tested, while requiring the same pumping

    power.

    2003 Elsevier Ltd. All rights reserved.Metal foams as compact high

    K. Boomsma a, D. Poa Laboratory of Thermodynamics in Emerging Technologies, Ins632-1176.

    E-mail address: [email protected] (D. Poulikakos).

    0167-6636/$ - see front matter 2003 Elsevier Ltd. All rights reservdoi:10.1016/j.mechmat.2003.02.001erformance heat exchangers

    kakos a,*, F. Zwick b

    of Energy Technology, Swiss Federal Institute of Technology,

    03) 11611176

    www.elsevier.com/locate/mechmatview of the history of the study of uid dynamics

    through porous media can be found in a review by

    ed.

  • of M

    ARTICLE IN PRESSNomenclature

    A area [m2]C form coecient [m1]CTE coecient of thermal expansion

    [mm1 K1]D diameter [m]FS Full scale

    K permeability [m2]L length [m]M compression factor []Nu Nusselt number (dened in Eq. (12)) []P pressure [bar]Q volumetric ow rate [m3 s1]R thermal resistance (dened in Eq. (12))

    [KW1]

    1162 K. Boomsma et al. / MechanicsLage (1998). An end result of the nearly 150 year-old work in porous media is the widely accepted

    equation (Eq. (1)) which governs the pressure drop

    of a uid passing through a porous medium.

    DPL lKV qCV 2 1

    The term DP is the pressure drop across the me-dium, L is the length of the medium in the owdirection, l is the dynamic viscosity of the uid, Kis the permeability of the medium, V is the clear-channel (Darcian) velocity of the uid, q is thedensity of the uid, and C is the form coecient ofthe medium.

    The extension of the uid dynamic work in

    porous media in a channel to include simultaneous

    convective heat transfer can be traced back as far

    as Koh and Colony (1974); Koh and Stevens

    Re Reynolds number (dened in Eq. (4)) []

    T temperature [C, K]V velocity [m s1]_WW pumping power [W]c specic heat [kJ kg1 K1]f friction factor []h convection coecient [Wm2 K1]j Colburn factor (dened in Eq. (10)) []k thermal conductivity [Wm1 K1]_mm mass ow rate [kg s1]q heat rate [W]w absolute error []

    Greek symbols

    D dierence []a thermal diusivity [m2 s1]e porosity fraction (range 100P e > 0) [%]l dynamic viscosity [kgm1 s1]m kinematic viscosity [m2 s1]q density [kgm3]

    Subscripts

    c coolant

    con convection

    aterials 35 (2003) 11611176(1975). Their analysis considered a simple slugow velocity prole through the porous medium,

    but the cooling eects generated by the presence of

    a porous medium were shown. In the course of

    developing better analytical models for channel

    convection, Kaviany (1985) presented an analyti-

    cal solution of the transport equations based on

    the quadratic-extended Darcy ow model (Eq.

    (1)). Advancements continued in numerical workon forced convection through packed beds of

    spheres which included the notable numerical

    work performed by Poulikakos and Renken (1987)

    and the corresponding experimental investigation

    (Renken and Poulikakos, 1988). Many other

    models have been proposed to account for other

    variables, such as wall eects (Kaviany, 1985;

    Mehta and Hawley, 1969), variable porosity(Amiri and Vafai, 1994; Nield et al., 1999), and

    cs cross-sectional

    e eective

    f uid

    hyd hydraulic

    inlet inlet

    outlet outlet

    p perimeterpl plates solid

    th thermal

  • kos, 2002). However, in the case of open-cell metal

    K. Boomsma et al. / Mechanics of M

    ARTICLE IN PRESSeven non-Newtonian uids (Chen and Hadim,

    1998, 1999). The great majority of these studies

    that are covered by Kaviany (1995) consider

    spherical media, which possess porosities in the

    range e 0:30:6. This conguration is known asa bed of packed spheres, which models uid owthrough sediment and other granular systems.

    The question arises how far these concepts

    based on spherical media can be taken to include

    other types of porous media. One type of a non-

    spherical porous medium is open-cell metal foam.

    The structure of open-cell metal foams (Fig. 1)

    opens itself to a wide variety of possible applica-

    tions which include, but are not limited to, lightweight high strength structural applications,

    mechanical energy absorbers, lters, pneumatic

    silencers, containment matrices and burn rate

    enhancers for solid propellants, ow straighteners,

    catalytic reactors, and more recently, heat ex-

    changers. The open-cell metal foam structure has

    the desirable qualities of a well designed heat ex-

    changer, i.e. a high specic soliduid interfacesurface area, good thermally conducting solid

    phase, and a tortuous coolant ow path to pro-

    mote mixing. Depending on the particular open-

    cell metal foam conguration, its specic surface

    area varies between approximately 500 to over

    10,000 m2/m3 in compressed form (ERG, 1999).

    The metal matrix can be manufactured from a

    high thermally conducting solid such as aluminum(ks 200 Wm1 K1) or copper (ks 400Wm1 K1), which, merely by its presence in astatic uid, dramatically increases the overall ef-

    fective thermal conductivity of the uid-system

    (Calmidi and Mahajan, 1999). The overall eective

    thermal conductivity of the soliduid system (keff )can be most generally described by the porosity (e)and the conductivities of the uid and solid phasesby kf and ks, respectively (Kaviany, 1995).

    keff ekf 1 eks 2This increase in overall eective thermal conduc-

    tivity of the soliduid system to a level above that

    predicted by Eq. (2) was shown in the 2-D con-duction model by Calmidi and Mahajan (1999).

    Noting the inuence of the structure on the ther-

    mal conductivity of the metal foam matrix,Boomsma and Poulikakos (2001) developed anfoams, the numerical models have had limited

    success, and the experimentation to verify these

    models is restricted, particularly to the coolant,

    which is typically air. In cooling electronics which

    generate a large amount of excess heat, a liquid

    coolant is generally preferred over air because of

    the greater thermal conductivity and specic heat

    capacitance. In view of these requirements, exper-iments using a forced liquid coolant are needed not

    only to investigate the feasibility of using open-cell

    metal foams as heat exchangers, but also to provide

    a basis against which numerical models can be

    compared. The goal of this investigation is to

    provide an experimental study of the performance

    of open-cell aluminum foam heat exchangers in a

    forced convection ow arrangement using a liquidcoolant, which is deionized, degassed water.

    Comparisons with existing heat exchanger systems

    for the application of cooling of electronics are also

    provided.

    2. Experiment

    2.1. Apparatus

    The goal of the experiment was to measure the

    hydraulic and thermal performance of the open-

    cell aluminum foams when used as heat exchang-

    ers in a forced convection ow arrangement. The

    concept was to direct the coolant ow through a

    rectangular channel in which the aluminum foamheat exchanger is placed, occupying the entire

    cross-section of the channel. A heater was attachedimproved analytical heat conduction model based

    on the idealized 3-D unit cell of an open-cell metal

    foam.

    There currently exist analytical (du Plessis et al.,

    1994; Diedericks and du Plessis, 1997; Smit and duPlessis, 1999; Lu et al., 1998) and numerical models

    (Calmidi and Mahajan, 2000; Lage et al., 1996) for

    the uid ow and heat transfer in packed beds of

    spheres and extensive databanks of uid ow and

    heat transfer experiments used as verication of

    these models (Antohe et al., 1997; Lage et al., 1997;

    Lage and Antohe, 2000; Boomsma and Poulika-

    aterials 35 (2003) 11611176 1163to the aluminum foam via the heat spreader plate,

  • Fig. 1. (a) Open-cell aluminum foam (T-6106 alloy) in its as-manufactured state with a porosity of e 92% and approximately 6 mmdiameter pores (ks 200 Wm1 K1). (b) Close-up view on an individual cell of the aluminum foam depicted in (a). (c) Open-cell metalfoam similar to that shown in (a), but after compression by a factor of four. (d) Close-up view of the compressed foam depicted in (c).

    Note the altered form of the individual cells that had originally resembled the cell in (b).

    1164 K. Boomsma et al. / Mechanics of Materials 35 (2003) 11611176

    ARTICLE IN PRESS

  • through which the heat was conducted and even-

    tually convected into the coolant stream. The

    characterization of the open-cell metal foam heat

    exchangers included measuring the temperature of

    the heater block, the temperature of the heat

    spreader plate, the coolant temperature at severallocations in the coolant ow, the power delivered

    to the heating device, and the pressure drop across

    the heat exchangers for various coolant ow rates.

    A general overview of the experimental apparatus

    is shown in Fig. 2. The channel assembly is shown

    in detail in two separate views in Fig. 3.

    Eight pressure taps measuring 0.2 mm in di-

    ameter were bored into the housing at variouslocations to measure the static pressure, as seen in

    Fig. 3(b). The outermost ports were used for the

    pressure drop calculations to avoid the static

    pressure variations generated by the acceleration

    and deceleration of the uid as it enters and leaves

    the metal foam. The other six ports were used as a

    symmetry check of the ow. During the experi-

    ments, the pressure variation between the left and

    right ports did not uctuate more than 3% and was

    therefore neglected.

    The metal foam heat exchanger housing was

    manufactured from Ryton R4, which has a lowthermal conductivity (0.3 Wm1 K1) and a rea-sonable CTE (22 106 m/m C). The static pres-sure drop of the coolant across the metal foam

    heat exchanger was measured by two dierent

    dierential pressure transducers corresponding to

    their individual pressure ranges. A Huba dieren-

    tial pressure transducer was used for measuring

    the pressure in the lower pressure range, from 0 to0.20 bar, with an accuracy of 0.5% FS (0.001

    bar). For the pressure range from 0.20 to 3.45 bar,

    an Omega (PX81DO-050DT) dierential pres-

    sure transducer was employed with an accuracy

    of 0.25% FS (0.009 bar). E-type thermocou-

    ples (chromel/constantan) of 0.15 mm diameter

    a Acqu

    Data Acquisition PC

    Coo

    Heate

    Power Supply Oscilloscope

    re th

    K. Boomsma et al. / Mechanics of Materials 35 (2003) 11611176 1165

    ARTICLE IN PRESSRotameter

    Dat

    Coolant Chiller/Recirculator

    Pressure RegulatingValve Thermocouples

    20.0

    Metal Foam

    Fig. 2. Schematic view of the experimental setup used to measuminum foam heat exchangers.Foam Housing

    Pressure Transducer

    USB isition Device

    Valve Array

    lant Flow Direction

    r Assembly

    e thermal and hydraulic characteristics of the compressed alu-

  • Mshows

    foam

    foam

    of M

    ARTICLE IN PRESSmeasured the temperatures at various locations of

    the experimental apparatus. The thermocouples

    were calibrated in a thermal bath to within 0.5 Cand were inserted through the 0.2 mm diameter

    pressure taps that were located in the bottom of

    the channel to measure the temperature of the

    coolant ow (Fig. 3(b)). The pressure drop mea-surements were performed both with and without

    the thermocouples inserted through the pressure

    taps to determine if their presence altered the

    pressure readings. No eects were observed.

    The data from the thermocouples and the

    pressure transducers were measured by a USB

    data acquisition device manufactured by IOTech.

    This device enabled the real-time measurement,recording, and display of the temperature and

    pressure drop data on the attached personal

    METAL FOA

    FLOWINLET

    FLOWOUTLET

    PRESSURE PORTS

    (a)

    HEATER ASSEMBLY

    Fig. 3. (a) Cross-sectional view of the foam test housing which

    assembly, and the path of the coolant ow. (b) Top view of the

    placement of the foam component of the assembled compressed

    1166 K. Boomsma et al. / Mechanicscomputer.

    The coolant was pumped through the experi-

    mental apparatus by a Neslab chiller. It provided a

    maximum coolant ow rate of approximately

    10 l/min, could dissipate a total of 1600 W, and

    regulated the coolant inlet temperature to within0.3 C.

    The coolant ow rate was measured using two

    dierent rotameters. The lower ow rate range

    varied from 0 to 1.0 l/min, which corresponded to

    a ow velocity of 00.21 m/s for a channel cross-

    section of 40.0 mm 2.0 mm. For this lower range,a Voogtlin rotameter with 1% FS (0.01 l/min) ac-curacy was utilized. The higher ow rate rangetested varied from 1.00 to 5.00 l/min, corre-

    sponding to a ow velocity range of 0.211.04 m/sfor a channel cross-section of 40.0 mm 2.0 mm.In the higher ow rate range, a Wisag 2000 ro-

    tameter was employed with an accuracy of 1% FS

    (0.11 l/min).

    The heating system consisted of a blockmachined

    from oxygen-free copper (ks 400 Wm1 K1)measuring 40.0 mm 44.0 mm 20.0 mm high. Fiveholes measuring 6.35 mm in diameter were bored

    through the block to hold ve 220 W cartridge

    heaters in place. The voltage and current delivered to

    the ve cartridge heaters were monitored by an os-

    cilloscope. The maximum power delivered to the

    heater block was 1100 W. Smaller holes measur-

    ing 0.2 mm in diameter were drilled into the top

    and base of the copper heater block, perpendicu-lar to the coolant ow direction. This allowed the

    insertion of small diameter thermocouples to mea-

    FLOWOUTLET

    FLOWINLET

    4cm7cm

    (b)

    the location of the aluminum foam heat exchanger, the heater

    test housing depicted in (a) with the lid removed to show the

    heat exchanger.

    aterials 35 (2003) 11611176sure the temperature dierence across the heater

    block.

    2.2. Aluminum foam heat exchangers

    One of the desirable qualities of the open-cellmetal foam in a heat exchanger application is the

    large specic surface area, which ranges from ap-

    proximately 500 to over 3000 m2/m3. Compressing

    the foam further increases this already large sur-

    face area to volume ratio. To generate an array of

    open-cell metal foam heat exchangers, 6101-T6

    aluminum alloy (ks 218 Wm1 K1) was castinto foam form at two dierent porosities ofe 92% and 95% with an average cell diameter of2.3 mm. This foam is listed as 40 PPI by the

  • phys

    76.0 82.5

    52.0 66.9

    Specic surface area [m2/m3] Measured porosity [%]

    2700 92.8

    50

    60

    70

    80

    90

    100

    0 2 4 6 8 10

    Expected Porosity (95%)Effective Porosity (95%)Expected Porosity (92%)Effective Porosity (92%)

    M

    [%]

    of M

    ARTICLE IN PRESSmanufacturer (ERG, 1999). The PPI acronym

    designates pores per linear inch, but due to the

    ambiguity of this label, the pore diameters of the

    uncompressed 40 PPI foam were visually measured

    by hand using a microscope and a scale calibrated

    to one-tenth of a millimeter and were tabulated in

    Table 1.The specic surface area of these foams was

    further increased by compressing them by a factor

    of M , which signies the ratio of the pre-com-pression to post-compression height of the foam

    block. In the one-dimensional compression pro-

    cess, the lateral sides of the metal foam are not

    Table 1

    Compressed foam physical data (Panel A), uncompressed foam

    Foam Compression Name

    Panel A

    5% 2 95-02

    4 95-04

    6 95-06

    8 95-08

    8% 2 92-02

    3 92-03

    6 92-06

    Panel B

    Foam Pore diameter [mm]

    40 PPI 2.3

    K. Boomsma et al. / Mechanicsrestrained. This is done to prevent any mass ac-

    cumulation in the foam caused by lateral materialmovements during the compression process.

    However, any material which ows to the out-

    side of the compression device is lost in the ma-

    chining of the foam to its nal overall dimensions,

    thus the resulting porosity of the foam may be

    higher than what would be predicted by Eq. (3).

    ecompressed 1M1 euncompressed 3The porosities of the aluminum foams were cal-

    culated by weighing them and comparing the

    density to that of solid 6101-T6 aluminum alloy.

    The corresponding expected porosities were also

    calculated by Eq. (3), using the manufacturersstated initial, uncompressed porosity of euncompressedand the given nominal compression factor, M .Table 1 lists both the measured and expected po-ical data (Panel B)

    Expected porosity [%] Measured porosity [%]

    90.0 88.2

    80.0 80.5

    70.0 68.9

    60.0 60.8

    84.0 87.4

    aterials 35 (2003) 11611176 1167rosities, and Fig. 4 shows their relationship

    graphically. The various congurations of alumi-

    num foams are named using two pairs of digits.

    The rst pair is pre-compression porosity; the

    number 92 designates euncompressed 92%. The sec-ond pair of digits designates the compression fac-

    tor, M . The foam 92-04, for example, was 92%porous in its uncompressed state and then com-pressed by a factor of four. In Fig. 4, the porosities

    of the 95% pre-compression foam samples that

    Fig. 4. Plot of the porosity of the raw foam material which was

    used in the aluminum foam heat exchangers. The lines denote

    the porosity predicted by the manufacturers stated initial po-rosity (e) and compression factor (M) based on Eq. (3), whilethe individual points are the measured values for the raw foam

    material.

  • given in length normalized units of [barm1] based

    of M

    ARTICLE IN PRESSwere measured closely follow the relationship as

    predicted by Eq. (3). However, the porosities of

    the 92% pre-compression foam remained higher

    than those predicted by Eq. (3) and the manufac-

    turers given data. By the consistency of the errorof the porosity data points from the predicted line

    for the 92% initial porosity foam, it can be as-

    sumed that this discrepancy is due to an inaccurate

    estimate of the initial porosity given by the foam

    manufacturer.

    The open-cell aluminum foams measured 40.0

    mm 40.0 mm 2.0 mm after the nal machiningprocess. To make them functional heat exchang-ers, each foam piece was brazed in a central posi-

    tion to an adjoining heat spreader plate which

    enabled a copper heating block to be mounted

    on the opposing side. Each heat spreader plate

    consisted of 6092 aluminum alloy with 18% SiC

    particles to increase the thermal conductivity to

    approximately 250 Wm1 K1 and measured 58.0mm 58.0 mm 1.9 mm thick.

    2.3. Experimental procedure

    Each open-cell aluminum foam heat exchanger

    was tested three times following the identical

    procedure. The heat exchanger was mounted into

    the test housing. The coolant was then pumped

    through the foam at the maximum attainable owrate, which varied according to the overall ow

    resistance of the individual heat exchanger. The

    inlet temperature of the coolant was held at 220.3

    C. After the 20 min stabilization period, fullpower to the heater cartridges was turned on, and

    the entire experimental apparatus was allowed to

    reach steady-state. Starting from the maximum

    ow rate, the temperatures at various locationswere measured and recorded in real-time via the

    USB data acquisition device and PC, which also

    recorded the pressure drop reading from the

    pressure transducer. The maximum coolant tem-

    perature allowed during operation was 100 C, atwhich the coolant began to vaporize.

    The ow rate was read from the rotameter.

    After the data were taken, the ow rate was re-duced, and the apparatus was given 5 min to reach

    steady-state for the next data point measurement.

    1168 K. Boomsma et al. / MechanicsAs noted in the work by Boomsma and Poulikakoson the 40.0 mm length of the heat exchangers, and

    the right-hand ordinate is the actual pressure drop

    measured in the experiments, and is given in the

    units of [bar]. As expected, those foams which

    possess the highest solid fraction (lowest e) as seenin Table 1 generated the largest pressure drop.(2002) and Antohe et al. (1997), the direction in

    which the ow rate is adjusted does not have an

    eect on the calculated permeability and form

    coecient needed to describe the ow resistance of

    a porous medium.

    3. Results and discussion

    3.1. Pressure drop

    The amount of work required to pump the

    coolant through a heat exchanger is a critical heat

    exchanger design parameter. In the work byBoomsma and Poulikakos (2002), the open-cell

    metal foams which comprised the in-ow compo-

    nent of the heat exchangers in this study were

    tested for their hydraulic characteristics. The pa-

    rameters used to describe the pressure drop char-

    acteristics of the foam heat exchangers are the

    permeability (K) and the form coecient (C)which are dened in Eq. (1). The metal foamcongurations used in this heat transfer study were

    produced under conditions identical to those used

    for the foams in the hydraulic characterization

    study of open-cell metal foams conducted by

    Boomsma and Poulikakos (2002). However, in

    assembling the heat exchangers, some of the alu-

    minum brazing material which attaches the foam

    to the heat spreader plate partially lled the poresat the interface which reduced the eective ow

    cross-sectional area of the heat exchanger and in-

    creased the ow resistance of the metal foam heat

    exchangers when compared to the results of the

    similar, unbrazed foams in Boomsma and Pouli-

    kakos (2002).

    The assembled heat exchangers were tested

    anew for their hydraulic characteristics and theresults of the pressure drop tests were plotted

    graphically in Fig. 5. The left-hand ordinate is

    aterials 35 (2003) 11611176These were led by the two most compressed foams,

  • brazed counterpart due to the presence of the

    brazing material in the pores of the open-cell metal

    foam at the brazing interface. Foams 95-08 and

    92-06 showed a slight decrease in the ow resis-

    tance when compared to their unbrazed counter-

    parts. This change in behavior by the two mosthighly compressed brazed foams can be attributed

    to warpage and distortion in the foam from the

    brazing process, thereby allowing ow bypass.

    With the remaining ve aluminum foam heat ex-

    changers, the amount of the increase of the ow

    resistance was not consistent. The change in the

    0

    10

    20

    30

    40

    50

    60

    70

    80

    0.0

    0.4

    0.8

    1.2

    1.6

    2.0

    2.4

    2.8

    3.2

    0.0 0.5 1.0 1.5 2.0

    95-0295-0495-0695-0892-0292-0392-06

    P/L[barm-1] P[bar]

    V [ms-1]

    K. Boomsma et al. / Mechanics of Materials 35 (2003) 11611176 1169

    ARTICLE IN PRESS95-06 and 95-08, with foam 92-06 generating

    nearly the same pressure drop. The foam which

    produced the lowest pressure drop was foam 95-02, which was also the most porous of the samples.

    The hydraulic characteristics of the brazed

    foams were calculated by using a least squares

    curve tting approach as described in Antohe et al.

    (1997) and Boomsma and Poulikakos (2002) to

    solve for the K and C in Eq. (1). These perme-ability and form coecient values were compared

    to the values obtained in Boomsma and Poulika-kos (2002), which used the same aluminum foam

    congurations, but without any brazing material.

    Table 2 compares the permeability and form co-

    ecients between the assembled heat exchangers

    Fig. 5. Pressure-drop curves for the metal foam heat ex-

    changers plotted on a length-normalized (DPL1) and actualpressure scale (DP ) against the Darcian ow velocity (V ).and the unbrazed foam blocks, along with the

    corresponding uncertainty percentages. Almost

    every assembled heat exchanger showed an in-

    crease in the ow resistance compared to its un-

    Table 2

    Flow resistance comparison

    Foam Unbrazeda

    K [1010 m2] C [m1] K [1010 m

    95-02 44.4 1168 34.4

    95-04 19.7 2707 6.87

    95-06 5.25 4728 3.16

    95-08 2.46 8701 2.52

    92-02 36.7 1142 30.8

    92-03 23.0 1785 8.26

    92-06 3.88 5518 3.95

    aResults reported in Boomsma and Poulikakos (2002).ow restriction depends upon the non-standard

    individual production process of each heat ex-changer.

    For a more general base of comparison, the

    hydraulic characteristics of the heat exchangers

    can be viewed using non-dimensional ow factors.

    One of which is the Reynolds number (Re) as de-ned for porous media in Kaviany (1995). For a

    degree of uniformity in the eld of porous media,

    the characteristic length used in the Re is replacedby the square root of the permeability (K) asshown in Eq. (4), where q is the density of theuid, V is the Darcian ow velocity, and l is thedynamic viscosity of the uid.

    Re qVK

    p

    l4

    The other commonly used non-dimensional

    ow describing factor is the Fanning friction fac-

    tor (f ) which is given in Eq. (5). This providesinformation concerning the required pressure drop

    (DP ) across a heat exchanger and comes into use

    Brazed heat exchanger

    2] rk [%] C [m1] rC [%]

    13.4 1276 2.9

    7.0 2957 3.2

    8.7 5066 9.8

    7.4 4731 5.6

    10.2 1472 5.0

    9.1 2820 3.9

    6.7 3399 5.4

  • of Materials 35 (2003) 11611176

    ARTICLE IN PRESSwhen the heat transfer performance-to-cost ratio is

    considered.

    f DP4 LDhyd

    qV 2

    2

    5In Eq. (5), the hydraulic diameter (Dhyd) is de-scribed by Eq. (6)

    D 4Acs 6

    Re

    f

    1

    10

    100

    0 20 40 60 80 100 120 140

    95-0295-0495-0695-0892-0292-0392-06

    Fig. 6. The calculated friction factor (f ) of the aluminum foamheat exchangers based on Eq. (5) plotted against the Re as de-ned in Eq. (4).

    1170 K. Boomsma et al. / Mechanicshyd Lp

    with Acs being the cross-sectional area of the owchannel (80.0 mm2), and Lp being the wetted pe-rimeter of the ow channel (84.0 mm). Fig. 6 plotsthe friction factor (f ) of Eq. (5) against the Re, ascalculated in Eq. (4). The plot of the friction factor

    levels o after a permeability based Re of ap-proximately 20. In this range, the pressure drop

    over the foam is dominated by the form coecient

    of Eq. (1). This is in agreement with the published

    results on the pressure drop of the ow through

    both uncompressed and compressed metal foamsin Boomsma and Poulikakos (2002).

    3.2. Heat exchanger performance

    A practical measure of the performance of a

    heat exchanging device is the dimensionless Nus-

    selt number (Nu) (Bejan, 1995) as given in Eq. (7).Nu hDhydkc

    qAconDT

    Dhydkc

    7

    The coecient h is the convection heat transfercoecient which characterizes the heat transfer

    between a solid and a uid. The thermal conduc-

    tivity of the coolant is given as kc, and the ow ofheat driven by a temperature dierence of DT isrepresented as q. Dhyd is the hydraulic diameter asgiven in Eq. (6). For uniformity in comparingthese results to the those from other investigations,

    the convection surface area, Acon, was consideredto be the interface area between the open-cell

    aluminum foam and the heat spreader plate (1600

    mm2).

    The temperature reference in Eq. (7) is an ar-

    bitrary variable, because the location of the refer-

    ence temperature positions is subjective. It was notpossible to reliably measure the temperature of the

    aluminum foam directly, so two small holes were

    drilled 1 mm deep into the top surface of the 1.9

    mm thick heat spreader plate on opposing sides of

    the copper heating block at central location, giving

    a temperature reference (Tpl) that was independentof the quality of the soldering between the heating

    element and the heat exchanger. It gave a goodmeasure of the average plate temperature when

    compared to the temperature dierence across the

    copper heater block, which experienced a maxi-

    mum temperature dierence from approximately

    18 C at the lower coolant ow speed range to lessthan 1 C at the upper end of the coolant owspeed range (V > 1:0 m/s). The other referencetemperature (Tc;inlet) was the temperature of thecoolant at the channel entrance. It was set to 295

    K and did not vary more than 0.3 K during the

    experimentation. The following is the nal relation

    for DT used in Eq. (7)

    DT Tpl Tc;inlet 8

    where Tpl is the temperature of the AlSiC heatspreader plate and Tc;inlet is the temperature of thecoolant at the inlet of the metal foam heat ex-

    changer. The heat transfer rate to the coolant (q) isdened by the following energy balanceq _mmcTc;outlet Tc;inlet 9

  • where _mm is the mass ow rate of the coolantpassing through the heat exchanger, c is the spe-cic heat of the incompressible coolant, and Tc;outletis the coolant temperature at the metal heat ex-

    changer outlet. The inlet and outlet coolant tem-peratures Tc;inlet; Tc;outlet were measured at 1.5 cmbefore and after the heat exchanger at the middle

    of the 2 mm channel height to accurately measure

    the mean temperature of the stream. The heat

    transfer rate which was evaluated with Eq. (9) was

    then back checked against the measurement of the

    power deliver to the heater cartridges to ensure

    that the mean temperature of the coolant streamwas being measured. With the substitution of Eqs.

    (8) and (9) into Eq. (7), the Nu was calculated bythe following expression from the experimental

    K. Boomsma et al. / Mechanics of M

    ARTICLE IN PRESSdata.

    Nu qAconTpl Tc;inlet

    Dhydkc

    _mmcTc;outlet Tc;inletAconTpl Tc;inlet

    Dhydkc

    10

    The Nusselt numbers for the open-cell aluminum

    foam heat exchangers were calculated at various

    coolant ow rates and plotted against the coolant

    ow speed in Fig. 7. The bare AlSiC heat spreader

    plate is also included in this comparison, and is

    labeled as plate.All Nusselt numbers began at zero for a zero

    coolant ow velocity and increased monotonically

    0

    20

    40

    60

    80

    100

    120

    140

    0.0 0.5 1.0 1.5 2.0

    95-0295-0495-0695-0892-0292-0392-06plate

    V [m/s]

    Nu[-]

    Fig. 7. The heat convection quantifying Nu as calculated in Eq.

    (7) plotted against the Darcian ow velocity (V ).with increasing coolant velocity. In the lower

    coolant ow velocity range, up to 0.729 m/s, the

    aluminum foam heat exchanger 92-06 achieved the

    largest Nu values. At the coolant ow velocityvalue of 0.729 m/s, the Nu of foam 95-04 surpassedthat of foam 92-06, and then continued for amaximum Nu of 134.6 at a coolant ow velocity of1.33 m/s. The foams with the lowest Nu valueswere the two most porous foams, 92-02 and 95-02,

    as seen by the comparison between Fig. 7 and

    Table 1, in which the porosities of the raw metal

    foam used for the heat exchangers are given. The

    Nu values of the remaining three aluminum foamheat exchangers, 92-03, 95-06, and 95-08, werenearly identical, except in the coolant velocity

    range under 0.50 m/s, where foam 95-08 had

    consistently a lower Nu value than the other twofoams, 92-03 and 95-06.

    Heat exchangers are commonly characterized

    by the Colburn j factor, which gives a heat transferperformance estimate comparing the convection

    coecient to the required ow rate of a heat ex-changer. This relationship is closely related to the

    friction factor in Fig. 6. The Colburn j factor isbased on the measured convection coecient (h),the necessary velocity of the coolant in order to

    achieve the corresponding convection coecient

    (V ), and the uids mechanical and thermal prop-erties, as described by the density (q), specic heat(c), kinematic viscosity (m l q1), and the uidthermal diusivity (a k q1 c1). The Colburnj factor is given in Eq. (11).

    j hqcV

    va

    2=311

    Fig. 8 plots the Colburn j factor against the Re,as dened in Eq. (4) and done in the work by

    Kays and London (1984), which has become one

    of the standard methods for reporting the per-

    formance of heat exchanging devices. Note that

    the bare plate is not included on this graph be-

    cause the Re in a clear channel is calculated in acompletely dierent manner from the metal foam

    experiments. The characteristic length in the Reof a clear channel is based on the hydraulic dia-

    meter (Dhyd), while in a porous medium, thecharacteristic length is derived from the perme-

    aterials 35 (2003) 11611176 1171ability (K).

  • 01

    2

    3

    4

    5

    0 5 10 15 20

    95-0295-0495-0695-0892-0292-0392-06

    W[W]

    (b)

    [K .kW 1]RthFig. 9. (a) Plot of the required pumping power ( _WW ) dened inEq. (12) against the corresponding thermal resistances (R ) as

    of M

    ARTICLE IN PRESS3.3. Pumping power considerations

    In any heat exchanger design, the heat convec-

    tion performance of the heat exchanger must beweighed against the energy required to operate the

    system, which is the pumping power in this con-

    guration. The required pumping power was cal-

    culated for the aluminum foam heat exchangers at

    various coolant ow velocities according to Eq.

    (12).

    Re

    j

    0.00

    0.05

    0.10

    0.15

    0.20

    0.25

    0 20 40 60 80 100 120 140

    95-0295-0495-0695-0892-0292-0392-06

    Fig. 8. The commonly used heat exchanger characterization

    parameter, the Colburn j factor of Eq. (11) plotted against theRe as dened in Eq. (4).

    1172 K. Boomsma et al. / Mechanics_WW DPQ 12In Eq. (12), _WW is the pumping power, DP is thepressure drop across the aluminum foam heat ex-

    changer, and Q is the volumetric ow rate of thecoolant passing through the heat exchanger.

    In addition to the Nu and the Colburn j factor,a common means to measure the heat convectioneectiveness is the thermal resistance. Lower

    thermal resistance facilitates the heat ow through

    the heat exchanger. Eq. (13) gives the common

    denition (Rth) for the thermal resistance in a heatconvection arrangement.

    Rth DTq Tpl Tc;inlet

    _mmcTc;outlet Tc;inlet 13

    The measured value of the power delivered to the

    heating cartridges via the oscilloscope was used as

    a back check for reasonable gures obtained by

    measuring the temperature dierence of the cool-0.001

    0.01

    0.1

    1

    10

    100

    0 50 100 150 200

    95-0295-0495-0695-0892-0292-0392-06plate

    [K .kW 1]Rth

    W[W]

    (a)

    aterials 35 (2003) 11611176ant across the heat exchanger. During the experi-

    ments, the power losses to the environment did not

    exceed 10%.

    The thermal resistances were calculated for the

    aluminum foam heat exchangers and the bare

    plate at various coolant ow velocities and were

    plotted in Fig. 9 against the required pumping

    power as dened in Eq. (12). Fig. 9(a) is a conve-nient log plot of the data for a general overview. In

    the corresponding plot of the same data in Fig.

    9(b) on a linear scale, the optimal design is that

    which minimizes the distance from the point to the

    origin of the plot. This point was obtained by

    foam 92-06, with a thermal resistance of 8.00

    th

    calculated in Eq. (13). (b) Close-up view of the pumping power

    versus thermal resistance plot of (a) showing foam 92-06 best

    approaching the ideal zero-approaching value for both the

    pumping power and thermal resistance.

  • KkW1 and a required pumping power of 1.29 Wat a coolant ow velocity of 0.356 m/s. The two

    other foams which also rated well in this perfor-

    mance to eciency comparison were foams 95-04

    and 92-03. The worst performance by a metal

    foam heat exchanger was generated by 95-08,

    eight individual heat exchangers. Therefore, the

    pumping power and thermal resistance results re-

    ported in Fig. 9 had to be further adjusted to es-

    timate the performance of a heat exchanger array

    consisting of eight individual metal foam heat ex-

    changers. To make these coolant and congura-

    al co

    [Wm

    K. Boomsma et al. / Mechanics of Materials 35 (2003) 11611176 1173

    ARTICLE IN PRESSwhich did not even t into the scale of Fig. 9(b). Its

    relatively poor performance can be attributed to

    the brazing process which also caused the unusu-

    ally low ow resistance for such a compressed

    foam (Fig. 5). The bare plate had the overall

    highest average thermal resistance. Comparing the

    bare plate against two foams at a relatively high

    thermal resistance of 50 KkW1, namely foams95-02 and 92-02, the plate required a pumping

    power which was 10 times greater than that re-

    quired by the two aforementioned foams.

    3.4. Heat exchanger comparison

    To evaluate their practicality as heat exchangers

    in industrial applications, the results of the heattransfer experiments performed on the aluminum

    foam heat exchangers were compared to test re-

    sults from an internal investigation performed by

    Asea Brown Boveri (Tute, 1998) on various heat

    exchangers. This evaluation was carried out by

    comparing the required coolant pumping power

    against the thermal resistance. However, since

    these readily available heat exchanger congura-tions were tested with a 50% waterethylene glycol

    solution, the aluminum foam heat exchanger ex-

    periments had to be scaled to account for the

    higher viscosity and lower thermal capacitance of

    the 50% waterethylene glycol solution. The rele-

    vant physical properties of water (Incropera and

    De Witt, 1990) and the 50% ethylene glycolwater

    coolant (Tute, 1998) are listed in Table 3.The experiments reported in Tute (1998) were

    conducted on a cooling system which consisted of

    Table 3

    Coolant properties at 300 K

    Property coolant Density

    q [kgm3]Therm

    k 103Water 997 61350% waterethylene glycol 1034 420tion adjustments, an approach based on the

    following assumptions was used to scale the per-

    formance of aluminum foam heat exchangers as if

    they were tested with a 50% waterethylene glycol

    solution in an array of eight individual heat ex-

    changers.

    1. The heat rate (q) for all varying ow conditionsremained unchanged from the water experi-

    ments to the 50% waterethylene glycol experi-

    ments.

    2. The operating temperatures measured in the

    water experiments remained unchanged.

    3. The volumetric ow rate of 50% waterethylene

    glycol solution was adjusted in proportion to its

    lower heat capacitance value and its higher den-sity to maintain the same heat transfer rate

    achieved in the experiments conducted with wa-

    ter. This adjustment required a 31% greater vol-

    umetric ow rate of the 50% waterethylene

    glycol mixture to achieve the same cooling ca-

    pacity under the rst assumption.

    4. To estimate the pressure drop across the alumi-

    num foam heat exchangers using the 50%waterethylene glycol solution, the permeability

    (K) and form coecient (C) obtained from thepressure drop experimentation were used in

    Eq. (1).

    5. The required pumping power for a cooling unit

    of eight heat exchangers is obtained by multi-

    plying the required pumping power of an indi-

    vidual heat exchanger by a factor of eight.6. The thermal resistance of a cooling unit consisting

    of eight individual metal foam heat exchangers

    nductivity1 K1]

    Dynamic viscosity

    l 103 [N sm2]Specic heat

    c [J kg1 K1]

    0.86 41793.19 3302

  • turbulence enhancers, and a simple at plate with

    a 0.2 mm channel height. In the experiments done

    by ABB (Tute, 1998), each individual heat ex-

    changer had a convection surface area in contact

    with the coolant measuring 34 mm in the length of

    the ow stream and 45 mm across, which providedan overall convective area of 1530 mm2. This is

    compared to the conguration for the experiments

    completed on the compressed aluminum foam heat20.1

    1

    10

    100

    100095-0295-0495-0695-0892-0292-0392-06plate

    W[W]

    Heat exch. with protrusions

    0.2 mm Gap

    1174 K. Boomsma et al. / Mechanics of Materials 35 (2003) 11611176

    ARTICLE IN PRESSwas estimated by dividing the thermal resistance

    of a single 1600 mm2 metal foam heat exchangerby a factor of eight.

    Fig. 10 plots the required pumping power (Eq.

    0.010 2 4 6 8 10 12

    ][ 1.kWKRth

    Brand name heat exch.

    Fig. 10. Plot of the predicted pumping power ( _WW ) versusthermal resistance (Rth) for the compressed aluminum foamheat exchangers when using a 50% waterethylene glycol

    coolant. These results are compared to the results of test on

    three commercially available heat exchanger congurations as

    reported in Tute (1998).(12)) against the thermal resistance dened in Eq.

    (13) for various compressed aluminum foam heat

    exchangers and the bare AlSiC plate as they would

    perform in an array of eight individual heat ex-

    changers using a 50% waterethylene glycol solu-tion. These results are overlayed onto the test

    results by ABB (Tute, 1998), in which three dif-

    ferent heat exchanger congurations were consid-

    ered. These tested heat exchangers consisted of a

    brand name heat exchanger, a generic at plate

    heat exchanger with small protrusions to act as

    Table 4

    Average uncertainty of coecients (%)

    Foam q Nu ReK

    95-02 17.9 19.3 11.8

    95-04 13.1 14.5 9.6

    95-06 14.5 25.8 11.9

    95-08 13.6 14.3 11.5

    92-02 16.3 16.9 13.2

    92-03 13.7 14.7 10.3

    92-06 12.6 15.1 10.3exchangers, which was 1600 mm , or 4.6% larger.

    The thermal resistances of the commercially

    available heat exchanger congurations in Tute

    (1998) were calculated in a manner identical to Eq.

    (13).Several observations can be made by scaling the

    water experiments to using the 50% waterethyl-

    ene glycol solution. Comparing the compressed

    aluminum foam heat exchangers in Fig. 10, it is

    clear that they generated an Rth that was lowerthan the best heat exchanger tested by ABB (Tute,

    1998) by a factor of nearly two. Another relevant

    observation is the preservation of the order of thepumping power curves of the various compressed

    aluminum foam heat exchangers. This means that

    the relative performance of the heat exchangers

    with the 50% waterethylene glycol solution can be

    well evaluated by using water. The only exception

    was the at AlSiC plate. Due to its high perme-

    ability (K) and correspondingly low form coe-cient (C), the pumping power requirement for theAlSiC plate increased by a disproportionately

    smaller factor than the compressed aluminum

    foam heat exchangers, and in Fig. 10, the perfor-

    mance of the AlSiC plate surpassed that of the

    worst performing compressed aluminum foam

    heat exchanger, 95-08, due to its relatively high

    thermal resistance.

    f j _WW

    16.5 21.6 17.1

    16.5 17.3 8.8

    20.6 28.6 7.8

    20.6 18.0 8.5

    13.5 19.1 15.7

    16.5 17.5 9.918.3 18.2 7.9

  • of f , labeled as wf , which is a function of V and Pis as follows:

    signicant improvement in the eciency over

    K. Boomsma et al. / Mechanics of Materials 35 (2003) 11611176 1175

    ARTICLE IN PRESSseveral commercially available heat exchangers

    which operate under nearly identical conditions.The metal foam heat exchangers decreased the

    thermal resistance by nearly half when compared

    to currently used heat exchangers designed for the

    same application.

    Acknowledgements

    It is gratefully acknowledged that this research

    was supported jointly by the Swiss Commission

    for Technology and Innovation (CTI) through

    project no. 4150.2 and by the ABB Corporatewf ofoV

    wV

    2 of

    oPwP

    2s14

    The average percentage values of these coecients

    are tabulated in Table 4.

    4. Conclusion

    Open-cell aluminum foams were compressed byvarious factors and then fashioned into heat ex-

    changers intended for electronic cooling applica-

    tions which dissipate large amounts of heat.

    Various common heat exchanger evaluation

    methods were applied to the data assembled from

    the extensive heat transfer experiments, which in-

    cluded the hydraulic characterization, the heat

    transfer performance, and an eciency study todetermine the most ecient metal foam heat ex-

    changer conguration for a particular heat trans-

    fer conguration. It was seen that the compressed

    aluminum foams performed well not only in the

    heat transfer enhancement, but they also made a3.5. Uncertainty analysis

    The coecients plotted in Figs. 510 were an-

    alyzed for their associated uncertainties following

    the standard procedure outlined in Taylor (1997).An example to nd the absolute uncertainty valueResearch Center, Baden-Daattwil, Switzerland.References

    Amiri, A., Vafai, K., 1994. Analysis of dispersion eects and

    nonthermal equilibrium, non-darcian, variable porosity

    incompressible-ow through porous-media. International

    Journal of Heat and Mass Transfer 37, 939954.

    Antohe, B.V., Lage, J.L., Price, D.C., Weber, R.M., 1997.

    Experimental determination of permeability and inertia

    coecients of mechanically compressed aluminum porous

    matrices. Journal of Fluids EngineeringTransactions of

    the ASME 119, 404412.

    Bejan, A., 1995. Convection Heat Transfer. John Wiley & Sons,

    Inc., New York.

    Boomsma, K., Poulikakos, D., 2001. On the eective thermal

    conductivity of a three-dimensionally structured uid-satu-

    rated metal foam. International Journal of Heat and Mass

    Transfer 44, 827836.

    Boomsma, K., Poulikakos, D., 2002. The eects of compression

    and pore size variations on the liquid ow characteristics of

    metal foam. Journal of Fluids EngineeringTransactions

    of the ASME 124, 263272.

    Calmidi, V.V., Mahajan, R.L., 1999. The eective thermal

    conductivity of high porosity brous metal foams. Journal

    of Heat TransferTransactions of the ASME 121, 466

    471.

    Calmidi, V.V., Mahajan, R.L., 2000. Forced convection in high

    porosity metal foams. Journal of Heat TransferTransac-

    tions of the ASME 122, 557565.

    Chen, G., Hadim, H.A., 1998. Forced convection of a power-

    law uid in a porous channelnumerical solutions. Heat

    and Mass Transfer 34, 221228.

    Chen, G., Hadim, H.A., 1999. Forced convection of a power-

    law uid in a porous channelintegral solutions. Journal of

    Porous Media 2, 5969.

    Darcy, H., 1856. Les Fontaines Publiques de la ville de Dijon.

    Dalmont, Paris.

    Diedericks, G.P.J., du Plessis, J.P., 1997. Modelling of ow

    through homogeneous foams. Mathematical Engineering in

    Industry 6, 133154.

    du Plessis, P., Montillet, A., Comiti, J., Legrand, J., 1994.

    Pressure-drop prediction for ow-through high-porosity

    metallic foams. Chemical Engineering Science 49, 3545

    3553.

    Dupuit, J., 1863. Etudes Theoriques et Pratiques sur le

    Mouvement des Eaux. Dunod, Paris.

    ERG, 1999. Duocel Aluminum Foam Data Sheet. ERG

    Material and Aerospace, Oakland.

    Forchheimer, P., 1901. Wasserbewegung durch Boden. Z. Ver.

    Deutsch. Ing. 45, 17361741, pp. 17811788.

    Hazen, A., 1893. Some physical properties of sand and gravels

    with special reference to their use in ltration. Massachu-

    setts State Board of Health, Twenty-fourth Annual Report.

    Incropera, F.P., De Witt, D.P., 1990. Introduction to Heat

    Transfer, second ed. John Wiley & Sons, New York.

    Kaviany, M., 1985. Laminar-ow through a porous channel

    bounded by isothermal parallel plates. International Journalof Heat and Mass Transfer 28, 851858.

  • Kaviany, M., 1995. Principles of Heat Transfer in Porous

    Media, second ed. Springer-Verlag, New York.

    Kays, W.M., London, A.L., 1984. Compact Heat Exchangers,

    third ed. McGraw-Hill, New York.

    Koh, J.C.Y., Colony, R., 1974. Analysis of cooling eectiveness

    for porous material in a coolant passage. Journal of Heat

    TransferTransactions of the ASME 96, 324330.

    Koh, J.C.Y., Stevens, R.L., 1975. Enhancement of cooling ef-

    fectiveness by porous materials in coolant passages. Journal

    of Heat TransferTransactions of the ASME 97, 309311.

    Kruuger, E., 1918. Die Grundwasserbewegung. InternationaleMitteilungen fuur Bodenkunde 8, 105.

    Lage, J.L., 1998. The fundamental theory of ow through

    permeable media from darcy to turbulence. In: Ingham,

    D.B., Pop, I. (Eds.), Transport Phenomena in Porous

    Media. Elsevier Science, Ltd., Oxford, pp. 130.

    Lage, J.L., Antohe, B.V., 2000. Darcys experiments and thedeviation to nonlinear ow regime. Journal of Fluids

    EngineeringTransactions of the ASME 122, 619625.

    Lage, J.L., Weinert, A.K., Price, D.C., Weber, R.M., 1996.

    Numerical study of a low permeability microporous heat

    sink for cooling phased-array radar systems. International

    Journal of Heat and Mass Transfer 39, 36333647.

    Lage, J.L., Antohe, B.V., Nield, D.A., 1997. Two types of

    nonlinear pressure-drop versus ow-rate relation observed

    for saturated porous media. Journal of Fluids Engineer-

    ingTransactions of the ASME 119, 700706.

    Lu, T.J., Stone, H.A., Ashby, M.F., 1998. Heat transfer in

    open-cell metal foams. Acta Materialia 46, 36193635.

    Mehta, D., Hawley, M.C., 1969. Wall eect in packed columns.

    I & EC Process Design and Development 8, 280282.

    Nield, D.A., Porneala, D.C., Lage, J.L., 1999. A theoretical

    study, with experimental verication, of the temperature-

    dependent viscosity eect on the forced convection through

    a porous medium channel. Journal of Heat Transfer

    Transactions of the ASME 121, 500503.

    Poulikakos, D., Renken, K., 1987. Forced-convection in a

    channel lled with porous-medium, including the eects of

    ow inertia, variable porosity, and brinkman friction.

    Journal of Heat TransferTransactions of the ASME

    109, 880888.

    Renken, K.J., Poulikakos, D., 1988. Experiment and analysis of

    forced convective heat-transport in a packed-bed of spheres.

    International Journal of Heat and Mass Transfer 31, 1399

    1408.

    Smit, G.J.F., du Plessis, J.P., 1999. Modelling of non-Newto-

    nian purely viscous ow through isotropic high porosity

    synthetic foams. Chemical Engineering Science 54, 645

    654.

    Taylor, J.R., 1997. An Introduction to Error Analysis, second

    ed. University Science Books, Sausalito.

    Tute, A., 1998. Versuchsbericht fuur die thermischen Messungenan Kuuhlkoorpern ohne oberaachenvergroossende Massnah-

    men. Asea Brown Boveri, VB8033.doc.

    1176 K. Boomsma et al. / Mechanics of Materials 35 (2003) 11611176

    ARTICLE IN PRESS

    Metal foams as compact high performance heat exchangersIntroductionExperimentApparatusAluminum foam heat exchangersExperimental procedure

    Results and discussionPressure dropHeat exchanger performancePumping power considerationsHeat exchanger comparisonUncertainty analysis

    ConclusionAcknowledgementsReferences