Sunteți pe pagina 1din 8

International Journal of Heat and Mass Transfer 67 (2013) 877884

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

Fluid temperature measurements inside metal foam and comparison to


BrinkmanDarcy ow convection analysis
Nihad Dukhan , Muntadher A. Al-Rammahi, Ahmed S. Suleiman
Department of Mechanical Engineering, University of Detroit Mercy, 4001 W. McNichols Rd., Detroit, MI 48221, USA

a r t i c l e

i n f o

Article history:
Received 22 April 2013
Received in revised form 19 August 2013
Accepted 21 August 2013
Available online 21 September 2013
Keywords:
Experiment
Metal foam
Porous media
Fluid temperature
Brinkman
Convection

a b s t r a c t
Actual air temperatures were locally measured inside commercial aluminum foam cylinder heated at the
wall by a constant heat ux, and cooled by forced air ow. The specially-developed experimental technique for such measurements is described in detail, and is shown to produce reasonably good data. Air
speeds were in the Darcy regime. The permeability of the foam was directly determined from experimental pressure drop points that were obtained using the same experimental set-up. The experimental air
temperatures are compared to their analytical counterparts. The volume-averaged analytical formulation
employed the Darcy-extended Brinkman model for momentum, and the non-thermal-equilibrium
two-energy-equation model for the temperatures of the solid and the uid phases inside the foam. The
solution steps, which are not new, are summarized. A comparison shows good agreement between the
experimental and the analytical air temperatures, given the complexity of the foams morphology and
the rounding nature of the volume-averaging technique. However, the analysis seems to under-predict
the uid temperature over most of the cross section. The experimental technique can be used for validation of other analytical solutions, numerical models and heat-exchange engineering designs based on
metal foam and similar porous media.
2013 Elsevier Ltd. All rights reserved.

1. Introduction
Convection heat transfer from the inside surface of porous media (when the pores are open to uid ow) has a wide range of
applications such as heat exchangers and chemical reactors.
Open-pore metal and graphite foams are excellent candidates for
such designs. Among the important characteristics of these foams,
from a heat transfer point of view, are the relatively high conductivities of the solid phase and the large surface area per unit volume. This is in addition to the vigorous mixing of the owing
uid due to the internal structure of these materials, which enhances convection between the solid and the uid, and gives rise
to an added mechanism of heat transfer called dispersion.
The internal architecture of open-pore foams in general is complex and random. Exact solutions of the complete transport equations are virtually impossible [1,2]. Researchers have solved
simplied forms of the governing equations, and relied on numerical simulations. Calmidi and Mahajan [3] numerically studied
forced convection of air ow in aluminum foam. The solid and uid
temperatures decayed gradually as the distance from the heated
wall increased. Hwang et al. [4] indicated that the local Nusselt

Corresponding author. Tel.: +1 313 993 3285; fax: +1 313 993 1187.
E-mail address: nihad.dukhan@udmercy.edu (N. Dukhan).
0017-9310/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijheatmasstransfer.2013.08.055

number increased with increasing Reynolds number for aluminum


foam. Angirasa [5] numerically studied convection heat transfer
due to water ow in metal foam heat dissipaters. He invoked local
thermal equilibrium. Poulikakos and Renken [6] provided numerical results for a channel lled with a uid-saturated porous medium, also using local thermal equilibrium. Kim and Jang [7]
investigated the validity of the local thermal equilibrium assumption itself.
Lee and Vafai [8] presented an analytical model for the solid and
the uid temperatures for Darcy ow in porous media between
two parallel plates. They assumed local thermal equilibrium at
the heated base, i.e., the solid, uid and wall temperatures are all
equal. They identied three regimes dominated by uid conduction, solid conduction or convection between the solid and the
uid. Vafai and Kim [9] solved the local-thermal-equilibrium
governning equations for a porous medium between two heated
parallel plates.
Haji-Sheikh and Vafai [10] provided analysis of heat transfer in
porous media imbedded inside ducts of different shapes, by solving
the governing equations assuming local thermal equilibrium, and
applying a constant-wall-temperature boundary condition. Minkowycz and Haji-Sheikh [11] solved the local-thermal-equilibrium
equations for the case of parallel plates and circular porous passages including the effect of axial conduction.

878

N. Dukhan et al. / International Journal of Heat and Mass Transfer 67 (2013) 877884

Nomenclature
b1
Bi
c
d1
d2
d3
Da
h
k
K
q00
T
u
U
r
R
ro
z
Z

constant, Eq. (16)


cylindrical porous media Biot number hrr 2o =2ks
(dimensionless)
specic heat of uid (J kg1 K1)
constant, Eq. (16)
constant, Eq. (16)
constant, Eq. (16)
Darcy number K=r 2o (dimensionless)
convection heat transfer coefcient (W m2 K1)
effective thermal conductivity (W m1 K1)
permeability (m2)
heat ux (W m2)
temperature (K)
volume-averaged pore velocity (m s1)
non-dimensional velocity
radial coordinate (m)
non-dimensional radial coordinate
radius of porous medium (m)
axial coordinate along ow direction (m)
non-dimensional axial coordinate along ow direction

Lu et al. [12] analyzed the forced convection problem in a tube


lled with a porous medium subjected to constant wall heat ux.
The two-equation model, which relaxes the thermal equilibrium
assumption, was solved. A closed-form solution for the solid and
the uid temperatures was presented. They used their results to
study tubes lled with metal foams as heat exchangers. In a follow-up study, Zhao et al. [13] presented an analytical solution for
a tube-in-tube heat exchanger, for which the inner tube and the
annulus were lled with a porous medium (metal foam). Both of
these studies employed the Brinkman-extended Darcy momentum
model. The researches in these two studies utilized the solutions
for investigating the effect of various foam parameters in practical
heat-exchange designs employing metal foam.
Analytical solutions in porous media continue to be sought after
due to their utility in practical engineering design, identifying
trends of critical variables, parametric studies and for validating
numerical models, see for example Xu et al. [14], Qu et al. [15]
and Xu et al. [16]. Direct comparisons to experimental values of
key variables, e.g., the uid and solid temperatures inside the foam,
is lacking in [1416], and seem to be non-existent in the literature
concerning heat transfer in metal foam. Experimental verications,
when possible, add condence to analytical and numerical solutions. In a recent comprehensive review, Zhao [17] indicated that
there is a lack of reliable experimental heat transfer data for
open-cell metal foam.
In the existing experimental studies concerning heat transfer in
metal foam, researchers typically measure substrate (wall) temperature, surface (skin) temperature and/or the temperatures at
the inlet and outlet of the foam. From such measurements, average
thermal transport parameters are determined, e.g., surface heat
transfer coefcient and volumetric heat transfer coefcient. Often,
these coefcients are used in obtaining Nusselt number which
serves as a basis for comparing analytical and numerical results
to their experimental counterparts. Calmidi and Mahajan [3], for
example, measured the wall temperature of aluminum foam sample bounded by substrates and heated. For comparing to their
numerical results, they used the measured wall temperature to obtain the average surface heat transfer coefcient. Bhattacharya
et al. [18] used the same kind of temperature measurements to

Greek

e
c
k
h

q
r
x
w

porosity (dimensionless)
ratio of effective viscosity to actual uid viscosity = l/le
ratio of effective thermal conductivities = kf/ks
dimensionless temperature
density of uid (kg m3)
surface area per unit volume
of porous medium (m1)
p
dimensionless parameter = pc
=Da

dimensionless parameter = Bik 1=k

Subscripts
f
uid
m
mean value
s
solid
w
wall
1
free stream

obtain the effective conductivity of metal foam, while Bhattacharya and Mahajan [19] used similar measurements to assess the
overall thermal performance of a nned-metal-foam heat sink.
Boomsma et al. [20] and Kurbas and Celik [21] conducted similar
measurements and calculations to investigate the performance of
a metal-foam compact heat exchanger. Kim et al. [22] measured
the wall, bulk inlet and outlet temperatures to obtain the spaceaveraged heat transfer coefcient and the Nusselt number for air
ow in metal foam. Similar measurements and calculations were
conducted by Zhao et al. [23]. Tzeng and Jeng [24] measured the
skin, inlet and outlet temperature for air ow in metal foam channel with 90-degree turned ow. The average Nusselt number was
calculated based on these temperatures. Hetsroni et al. [25] measured the wall temperature of a volumetrically-heated metal foam
sample using an infrared camera, and used this temperature to calculate the average heat transfer coefcient. Hwang et al. [4] used
wall and exit temperature measurements for air ow through metal foam, followed by iteration to back calculate the interstitial heat
transfer coefcient inside the foam. Noh et al. [26] measured the
wall temperature for an annulus lled with aluminum foam in order to calculate the local surface heat transfer coefcient. Kim et al.
[27] measured the inlet, outlet and wall temperatures for an aluminum-foam n in a plate-n heat exchanger. They used these temperatures to assess the thermal performance of the n.
In the current study, direct measurements of the uid temperatures inside commercial aluminum foam using a specially-developed technique are presented. Such measurements have an
intrinsic value. To the knowledge of the present authors, the measuring technique and the experimental data are novel, and have
not been performed or published previously. The experimental
technique for measuring the uid temperature can be used for validation of analytical and numerical models of heat transfer in
meso-scale porous media and for assessing the performance of
heat-exchange engineering designs based on such media.
The results of the current study are compared to their analytical
counterparts. For completeness, the analytical results are
demonstrated by concisely presenting and explaining the steps
and solution of the volume-averaged thermal- non-equilibrium
governing equations for forced convection heat transfer in a

879

N. Dukhan et al. / International Journal of Heat and Mass Transfer 67 (2013) 877884

cylindrical porous media due to Brinkman-extended Darcy ow. It


must be noted here that the solution and its steps are not new, but
have been reported in [1215]
2. The volume-average formulation
Consider a cylindrical isotropic porous medium of radius ro
bounded by an impermeable wall as shown schematically in
Fig. 1. The system is heated at the wall by a constant heat ux
q00 . There is fully-developed one-dimensional ow of a Newtonian
uid in the z-direction with a volume-averaged pore velocity component u. Upon solving the Brinkman-Darcy momentum equation
inside and outside the porous boundary layer [9], the velocity can
be obtained [10,11,28] as:

U 1

Io xR
I o x

where R = r/ro Da K=r 2o (the Darcy number), K is the permeability,


c = l/le, l is the uid viscosity, le is the effective viscosity,
U = u/
p
u1, u1 is the velocity outside the boundary layer, x c=Da and
Io is the modied Bessel function of order zero.
For thermally fully-developed conditions, the volume-averaged
two-energy-equation model is



ks @
@T s
 hrT s  T f 0
r
r @r
@r



kf @
@T f
@T f
hrT s  T f q c u
r
r @r
@r
@z

@T s @T f

0
@r
@r

at y r o ks

@T f
@T s
kf
q00
@r
@r

and T s T f T w

where Tw is the wall temperature (a function of the ow direction),


which is not known a priori [8]. The equality of the temperatures at
the heated wall as described in Eq. (5) has been used in the literature [1214]. Eqs (2)(5) are made non-dimensional:

q
ro

Ts

Tf



@hf
k @
Bihs  hf 2U
R
R @R
@R

at R 0

@hs @hf

0
@R
@R

at R 1

@hf
@hs
k
0 and hs hf 0
@R
@R

where R = r/ro, Z = z/ro, hs T s  T w =q00 r o =ks ; hf T f  T w =q00 ro =ks ,


k kf =ks and Bi hrr 2o =ks is the Biot number. The derivative ohf/oZ,
which is constant for thermally fully-developed conditions, is obtained by integrating the sum of Eqs. (2) and (3) over the cross-sectional area and applying the boundary conditions. This integration
results in @hf =@Z 2q00 =qcum r o
3. Solution
The two energy Eqs. (6) and (7) are added, and U is substituted
for from Eq. (1) to yield:






1 @
@ 
Io xR
R
hs khf
2 1
R @R
@R
I o x

Fig. 1. Schematic of the porous-media heated cylinder.




1 2
2 Io xR
R 1  2
1
2
x Io x

11

Eqs. (6) and (7) are de-coupled by (a) solving Eq. (11) for hs in terms
of hf and (b) substituting for hs, ohf/oZ and U in Eq. (7). The following
equation is obtained after rearranging:

 

1 dhf Bik 1
1
2
1
Bi
 2 R2


Bi

h
f
2
2
R
k
k
2
2
dR
x
dR
 2


x  Bi Io xR
2
I o x
x2
2

d hf

12

which is a single non-homogenous ODE for the uid temperature.


The general solution of Eq. (12) is obtained by rearranging the
homogenous part to take the general form of Bessel equation
[35,36]. The particular solution to Eq. (12) is sought by assuming
an appropriate form (according to the forcing function, i.e., the
right-hand side of Eq. (12)). The complete solution of Eq. (12) is

hf b1 I0 wR d1 I0 xR d2 R2 d3
where

p
w Bik 1=k

and

the

13
constants

are

given

by

x
1
2k
2 d3
b1  d1 Io wd
, d1 x2 Io x2Bi
, d2 2k1
and d3 Bik1
2
Io w
kx2 Bik1
2

x
 2Bi1=22=
. In obtaining the homogenous solution, a complex
Bik1
p
root i Bik 1=k is encountered. When the root is found to be
imaginary, as the case here, Bessel function Jo is replaced by the
modied Bessel function Io in the solution [35,36], as was done in
[14,15] in which a cylindrical system related to current problem
was solved, and was not done in [12,13] when solving an identical
problem to the one in hand. More details regarding this issue are
available in [37].
The solid temperature can readily be obtained by substituting
for hf from Eq. (13) into Eq. (11), which gives

hs

10

which is a non-homogenous ordinary differential equation (ODE)


with the dependent variable hs khf . The solution is

hs khf

where ks and kf are the effective thermal conductivities, Ts and Tf are


the volume-averaged temperatures of the solid and the uid inside
the open-pore porous medium, respectively, h is the interstitial heat
transfer coefcient between the solid and the uid, r is the surface
area per unit volume of the porous medium and q and c are the density and the specic heat of the uid, respectively. The Pclet number is assumed to be high such that the longitudinal conduction for
both the solid and the uid are negligible. For inclusion of these
terms, see Minkowycz and Haji-Sheikh [11]. Equations (2) and (3)
are well known [3,8,2934]. The following boundary conditions
apply:

at r 0



1 @
@hs
 Bihs  hf 0
R
R @R
@R





1
2
 kd2 R2 
kd1 Io xR  kb1 I0 wR
2
2
x Io x


1
2

kd3

2 x2

14

880

N. Dukhan et al. / International Journal of Heat and Mass Transfer 67 (2013) 877884

4. Experiment
An experimental heat transfer model was designed, fabricated
and tested, in order to allow for direct measurement of the uid
temperature, and subsequently for comparing the measured values
to the predictions of the volume-averaged analytical solution. The
model was essentially a cylindrical aluminum tube lled and
brazed to an aluminum foam core. Brazing of the two similar materials minimized the thermal contact resistance. The tube had a
length of 15.24 cm (6.0 in) in the ow direction, an inside diameter
of 25.56 cm (10.1 in) and a tube thickness of 6.4 mm (0.25 in). The
foam was obtained commercially (ERG Materials and Aerospace
[38]), and was made from aluminum alloy 6101-T6. It was marked
by an approximate industrial designation as 20-ppi (pores per linear inch). The porosity of the foam was 91% (calculated from measurements of its volume and weight), while other geometric
parameters were based on manufacturers data.
To isolate and measure the uid temperature inside the foam
using common thermocouples, each thermocouple was shielded
in a specially designed small perforated aluminum tube having
an inner diameter of 3.15 mm and an outer diameter of 4.55 mm.
The tube was perforated with circular holes 3.8 mm in diameter,
Fig. 2(a). A thermocouple was inserted in each perforated tube
and xed in place using high-temperature epoxy, blocking as few
of the perforated tubes holes as possible. The bead of each thermocouple was positioned such that it did not extend out of the small
tube, and remained shielded and protected by the wall of the small
tube, as seen in Fig. 2(a). As such, when the tube and its thermocouple are inserted in the foam, the bead would not touch the solid
ligaments of the foam or the wall of the small tube.
Two sets of ten-holes were drilled through the wall of the aluminum cylinder and the foam core reaching pre-determined radial
distances inside the foam. The rst set was at a distance of 3.81 cm,
while the other set was at 6.35 cm form the foam entrance. The radial depth of the holes, measured from the outer surface of the
tube, were 2.03, 3.30, 4.57, 5.84, 7.11, 8.38 9.56, 10.92, 12.19 and
13.46 cm (0.8, 1.3, 1.8, 2.3, 2.8, 3.3, 3.8, 4.3, 4.8 and 5.3 in). The
holes were arranged around the cross section with an angle of
36o between each two adjacent holes (Fig. 3), and were organized
in order to minimize alteration of the internal structure of the foam
and the interference with air ow through the foam. The diameters
of these holes were slightly larger than the diameters of the small
perforated tubes to allow tight t. The small perforated tubes with

Fig. 3. Location and arrangement of thermocouple holes around the cross section;
numbers represent the depth of each hole in cm.

their thermocouples were inserted in these holes, as seen in Fig. 2


(b). This allowed for air temperature measurements at different
known radial distances inside the foam. The perforation made
the small tubes more compatible with the internal structure of
the foam and allowed air to ow through while minimizing blockage to the ow.
Slots were machined on the outer surface of the aluminum cylinder to house the thermocouple wires, Fig. 2(b). After the thermocouples in their perforated small tubes were inserted in the foam,
the lead wires of the thermocouples were placed in the especiallymachined slots and directed away. Thermal epoxy then lled the
remaining volume of the slots, to end up with a smooth surface
matching the original outer surface of the aluminum cylinder,
Fig. 2(c). This insured that there were no air pockets trapped in
the wall thickness. Other small shallow holes were drilled in the
wall of the aluminum cylinder to house thermocouple beads for
measuring the wall temperature at various locations in the ow
direction (not shown). On the outer surface of the aluminum cylinder, thermofoil heaters (made by Minco Products) were attached

Fig. 2. Experimental aluminum foam heat transfer model construction: (a) Thermocouple in its small peroferated aluminum tube (thermocouple assembly), (b) Themocouple
assemblies inserted into holes with the leads directed outward in their slot, (c) A slot lled with thermal epoxy; Two holes for pressure drop measurement and (d) Thermofoil
heaters attached to outer surface of the heat transfer model.

881

N. Dukhan et al. / International Journal of Heat and Mass Transfer 67 (2013) 877884

using thermal adhesive. The heaters covered the outside area,


Fig. 2(d). Each heater had a resistance of 22.3 X, and could provide
up to 645 W of power. The whole assembly was then covered with
a 2.54-cm-thick insulation sheet.
Experiments were performed in an open-loop wind tunnel as
shown in Fig. 4. A suction unit (SuperFlo 600 Bench), which could
produce air ow rates up to 17 m3/min (600 ft3/min) was used to
induct room air into the tunnel and through the foam sample.
The suction unit was powered by eight fans; they were electrically
modied so that two can be on at a time, which was suitable for
producing low ow rates through the tunnel. A variable ow controller provided further adjustment of the airow through the tunnel and the test section.
Two holes were drilled at the top of the test section at a distance
of 5.08 cm (2 in), Fig. 2(c). The diameters of these holes were
7.9 mm (0.31 in) and they housed pressure measurement tubing.
The pressure drop was measured using an Omega differential pressure transmitter with a range 0746 Pa (0 to 3 in H2O).
As mentioned above, the test section of the tunnel was formed
by brazing the foam sample to the inside surface of an aluminum
cylinder. A reducing nozzle connected the exit of the test section
to a ow-measurement section. The tunnel section that housed
the speed measurement device was circular with a diameter of
63.5 mm (2.5 in). As such, the measured air speed at the ow meter
had to be adjusted, using conservation of mass and the ratio of the
diameters of the test section and the speed measurement section,
in order to obtain the mean velocity in the test section (and
through the foam). The low velocities in the foam of the current
experiment were achieved due to the relatively large crosssectional area of the experimental model. For velocity measurement, an Extech gas ow meter that could measure speeds up to
35 m/s (29 ft/min) was used.
The test section was secured in place and sealed. The pressure
transducer and the ow meter were connected to a data acquisition system (DAQ by Omega Engineering) which delivered the
readings to a computer. The ow rate was then varied to realize
different velocities in the test section. For each velocity, the steady-state static pressure drop was measured using the pressure taps
and the differential pressure transducer. The measurements were
repeated three times and the average of the three runs was reported. Form these data, the permeability was determined, as will
be shown later.
The heat transfer experiment proceeded as follows. The free
leads of all thermocouples were connected to the data acquisition
system, where the temperatures could be displayed and recorded
using a computer. The volumetric ow rate of room air was adjusted using controls on the suction unit, such that the desired

speed was realized inside the foam. The surface heaters were powered, and the power input was adjusted using a variac, so that the
desired power, and hence the heat ux to the foam, was achieved.
The air temperatures inside the foam, as well as the wall temperature, were monitored until there were no noticeable variations
in their readings, indicating steady-state conditions, which took
about 40 min. The steady-state air and wall temperatures were
recorded, as was the ambient air temperature.
5. Uncertainty analysis
The uncertainty in the velocity measurement had a contribution
from a xed error, ef = 2% (provided by the manufacturer) and a
random estimated error, er = 10% in each reading. For the pressure
transducer ef was 5% and er was 7% (these were provided in a calibration certicate). The total uncertainties in the pressure and
velocity were calculated by the root-sum-squares method according to Figliola and Beasley [39]:

q
d  e2f e2r

15

This resulted in a total uncertainty in the pressure drop dDp of 8.6%,


and in the air velocity dum of 10.2%. The length and diameter of
each sample were measured using a caliber. The uncertainties in
these readings were dL = dD = 1.0 mm, or 1.9%.
Further uncertainty analysis was performed in order to obtain
the uncertainty in the computed values of the permeability K from
curve-tting of the experimental pressure drop and velocity data.
According to Darcys law, the pressure drop per unit length of
the porous medium is linear with the velocity, i.e.:

Dp
AV
L

16

where A is a curve-t constant, and K = l/A. The percent uncertainty


in K is a result of uncertainties in its constituents l and A, and is
given by [39]

s
 2  2
dl
dK
dA


K
l
A

17

where dl and dA are the uncertainties in l and A, respectively. The


uncertainty in l was estimated as 1  107 N s/m2, taken as the
accuracy of the reported values in property tables [40]. This value
is small enough to cause negligible impact on the overall uncertainty in K, therefore it was ignored. Therefore, the uncertainty in
K is simply given by:

dK
dA
  100%
K
A

18

The average uncertainty in A is the same as the uncertainty in


according to Eq. (16). So it is given by

s
 2  2  2
dA
dDP
dL
dV


A
Dp
L
V

Fig. 4. Photograph of the experimental set-up.

Dp
;
LV

19

Using a reference velocity of 0.2 m/s and a reference pressure drop


of 1.6 Pa, the uncertainty in A, which is the same as the uncertainty
in K, was obtained as 13.4%.
The uncertainty in the temperature measurement was based on
the accuracy of the thermocouple 0.2 C, provided by manufacturer. The effective thermal conductivity of the solid aluminum
ligaments of the foam was obtained from an analytical one-dimensional model given by Calmidi and Mahajan [41]. These researchers
indicated that this model was excellent in matching their
measured values of the effective conductivity. Actually, similar
foam (20 pores per inch, 90.6% porous foam, made by the same
manufacturer) was tested in [41], and an effective thermal

882

N. Dukhan et al. / International Journal of Heat and Mass Transfer 67 (2013) 877884

and is occurring at a Darcy velocity of about 0.3 m/s. For velocities


below 0.3 m/s, the ow regime is purely Darcian. So, for heat transfer measurements, a velocity of 0.2 m/s was chosen to insure that
the ow was well within the purely Darcy regime.
The Darcy-regime permeability was determined from the
Darcy-regime pressure-drop data points only, according to Eq.
(21), as was done by Dukhan and Ali [42]. The air viscosity, l,
was taken as 1.8  105 kg/m s.

Table 1
Parameters for the metal-foam lled model.
Parameter

Value

ro
um
q00

0.128 m
0.2 m/s
8494.5 W/m2
1313.8 m2/m3
1.16  1007 m2
6.62 W/m K
207.8 W/m2 K
682.9
340.8
303.1
0.0044

r
K
ks
h
Bi

x
w
k

6.2. Heat transfer results and comparison

conductivity was obtained as 6.9 W/m K which is only 4.0% different from the value obtained for the foam used in the current study,
as reported in Table 1, which is encouraging. The uncertainty in the
effective conductivity was conservatively assumed to be 10%. The
uncertainty in the heat ux was assumed to be 10%, while the
uncertainty in the radius was 1.9%. These values along with the difference between the uid temperature and the wall temperature at
the center of the channel (79.7 C), which is the maximum difference that would result in the highest uncertainty, were substituted
in the following equation

dhf

hf

s

2 
2  2  2  2
dT f
dT w
dks
dq00
dr o

Tw  Tf
Tw  Tf
ks
q00
ro
20

Eq. (20) yielded an uncertainty in the non-dimensional temperature


of the uid equal to 14.3%.
6. Results
6.1. Pressure drop and permeability
The Forchheimer-extended Darcy equation for the pressure
drop, after dividing by the Darcy velocity is

Dp l
qCV
LV K

21

where, as before, K is the permeability and C is a form drag coefcient. The measured pressure drop is plotted against the Darcy
velocity, according to Eq. (21), in Fig. 5. The Darcy regime is identied by the absence of the last term (the Forchheimer form drag
term) in Eq. (21). In the purely Darcy regime, Eq. (21) is a horizontal
line with a y-intercept equal to l/K. When the form drag is important, Eq. (21) applies and we get a line with a slope equal to qC. This
change from Darcy to Forchheimer regime is clearly shown in Fig. 5,

Fig. 5. Experimental pressure drop versus Darcy velocity.

For the sample of the current study, the parameters of Table 1


were obtained by direct measurement or by computation, as follows. The surface area density of the foam was calculated form a
correlation given by the manufacturer, as shown in [43]. The effective thermal conductivity of the solid, ks, was obtained from a model given in Calmidi and Mahajan [41]. The interstitial heat transfer
coefcient inside the foam, h, was computed based on a correlation
given by Kuwahara et al. [44].
Fig. 6 shows the measured wall temperature along the ow
direction for ve different locations: 1.27, 3.81, 6.35, 10.16 and
12.7 cm. This temperature increases in the ow direction as expected. There is a subtle but identiable change of slope occurring
at 6.35 cm from the entrance- a sign of change from thermally
developing to fully-developed region. This trend in the wall temperature of heated metal foam is very similar to what has been presented in Calmidi and Mahajan [3], with the change in slope
occurring at an axial distance of about 8.3 cm from the entrance.
This difference in the thermal entry length is most likely due to
the difference in air speeds: 0.2 m/s in the current study and
0.61 m/s in Calmidi and Mahajan [3] and the number of pores
per inch: 20 in the current study and 5 in [3]. In the current investigation, the experimental case for the axial location of 6.35 cm
was taken to be in the thermally fully-developed region, while
the case at 3.81 cm was taken as nearly thermally fully-developed,
as will be shown next.
Fig. 7 shows the non-dimensional uid temperature as a function of the radial distance R. The solution given by Eq. (13) is plotted using a solid line, while the experimental data points are given
using the black solid circles. Generally speaking, the two temperature curves decrease gradually as the distance from the heated wall
(R = 1) increases, as expected.
The two temperature curves are relatively close to each other,
with the analytical solution predicting a slightly lower temperature, in general. In addition to experimental errors, other errors
may have been introduced into the analytical solution. The solution required the computation of the heat transfer coefcient inside the foam. This was obtained from a correlation based on a
geometrical modeling of the foam structure, Kuwahara et al. [44].
Similarly, the effective thermal conductivities ks and kf were obtained from a correlation also based on geometrical modeling of
the foam [41]. One possible error may be the existence of a temperature slip condition at the wall for the uid phase. Sahraoui and
Kaviany [45] indicated that there are physical non-uniformities
near the solid wall that bounds a porous medium, which alters
the value of the effective conductivities of the solid and uid
phases Moreover, the nature of the wall boundary condition when
a constant heat ux is applied is still a topic of debate, i.e., how the
total heat ux is split between the solid and the uid phases, see
for example Vafai and Yang [46].
The two points closest to the center of the foam cylinder are
seen to produce lower non-dimensional uid temperature, or higher physical temperatures, noting that hf T f  T w =q00 r o =ks . This is
true for both of axial locations z = 3.81 and 6.35 cm. Inspection of
Fig. 3, which shows the arrangement of the thermocouples around
the cross section, indicates that these two points are marked by the

N. Dukhan et al. / International Journal of Heat and Mass Transfer 67 (2013) 877884

883

Fig. 6. Wall temperature as a function of axial distance in the ow direction.

validating complex analytical and numerical modeling of the heat


transfer phenomenon in open-cell meso-porous media.

References

Fig. 7. Experimental and analytical uid temperatures.

numbers 13.46 and 12.19 cm. The gure clearly shows that the
blockage caused by inserting the two small perforated tubes with
their thermocouples is signicant enough to cause the temperature
at these two locations to be higher than what the temperature
would have been if the structure of the foam was not disturbed.

7. Conclusion
Direct measurement of the uid temperature inside a heated
cylinder lled with metal foam was conducted using a speciallydesigned technique. Scrutiny of published literature on heat transfer in metal foam reveals that such measurement has been lacking.
The data points were obtained in the Darcy regime and in the thermally fully-developed region. In order to compare to analytical
predictions, the thermal-non-equilibrium two-equation model for
convection heat transfer in a heated cylinder lled with a porous
medium and heated at the wall was revisited. The solution steps
were summarized and the analytical uid temperature was presented. The solution employed the Brinkman-extended Darcy ow
momentum equation. Relatively good agreement between the analytical and experimental uid temperature was displayed. However, the analytical solution of the volume-averaged energy
equations seemed to under-predict the uid temperature over
most of the cross section. Nonetheless, the measurements give an
opportunity to look into the issue of how volume-averaged values
compare to physically-measured local values of common quantities such as the temperature inside porous media. The experimental technique can be of utility for heat transfer designs, and for

[1] K. Vafai, C.L. Tien, Boundary and inertia effects on convective mass transfer in
porous media, Int. J. Heat Mass Transfer 25 (8) (1982) 11831190.
[2] M.L. Hunt, C.L. Tien, Effect of thermal dispersion on forced convection in
brous media, Int. J. Heat Mass Transfer 31 (2) (1988) 301309.
[3] V.V. Calmidi, R.L. Mahajan, Forced convection in high porosity metal foams, J.
Heat Transfer 122 (2000) 557565.
[4] J.J. Hwang, G.J. Hwang, R.H. Yeh, C.H. Chao, Measurement of the interstitial
convection heat transfer and frictional drag for ow across metal foam, J. Heat
Transfer 124 (2002) 120129.
[5] D. Angirasa, Forced convective heat transfer in metallic brous materials, J.
Heat Transfer 124 (2002) 739745.
[6] D. Poulikakos, K. Renken, Forced convection in a channel lled with porous
medium, including the effect of ow inertia, variable porosity, and Brinkman
friction, J. Heat Transfer 109 (1987) 880888.
[7] S.J. Kim, S.P. Jang, Effects of the Darcy number, the Prandtl number and the
Reynolds number on the local thermal non-equilibrium, Int. J. Heat Mass
Transfer 45 (2002) 38853896.
[8] D.Y. Lee, K. Vafai, Analytical characterization and conceptual assessment of
solid and uid temperature differentials in porous media, Int. J. Heat Mass
Transfer 42 (1999) 423435.
[9] K. Vafai, S.J. Kim, Forced convection in a channel lled with a porous medium:
an exact solution, J. Heat Transfer 111 (1989) 11031106.
[10] A. Haji-Sheikh, K. Vafai, Analysis of ow and heat transfer in porous media
imbedded inside various-shaped ducts, Int. J. Heat Mass Transfer 47 (2004)
18891905.
[11] W.J. Minkowycz, A. Haji-Sheikh, Heat transfer in parallel plates and circular
porous passages with axial conduction, Int. J. Heat Mass Transfer 49 (1314)
(2006) 23812390.
[12] W. Lu, C.Y. Zhao, S.A. Tassou, Thermal analysis on metal-foam lled heat
exchangers. Part I: Metal-foam lled pipes, Int. J. Heat Mass Transfer 49 (15
16) (2006) 27512761.
[13] C.Y. Zhao, W. Lu, S.A. Tassou, Thermal analysis on metal-foam lled heat
exchangers. Part II: Tube heat exchangers, Int. J. Heat Mass Transfer 49 (1516)
(2006) 27622770.
[14] H.J. Xu, Z.G Qu, W.Q. Tao, Analytical solution of forced convection heat transfer
in tubes partially lled with metallic foam using the two-equation model, Int.
J. Heat Mass Transfer 54 (2011) 38383846.
[15] Z.G. Qu, H.J. Xu, W.Q. Tao, Fully developed forced convective heat transfer in an
annulus partially lled with metallic foams: an analytical solution, Int. J. Heat
Mass Transfer 55 (2012) 75087519.
[16] H.J. Xu, Z.G. Qu, W.Q. Tao, Thermal transport analysis in parallel-plate channel
lled with open-celled metallic foams, Int. J. Heat Mass Transfer 38 (2011)
868873.
[17] C.Y. Zhao, Review on thermal transport in high porosity cellular metal foams
with open cells, Int. J. Heat Mass Transfer 55 (2012) 36183632.
[18] A. Bhattacharya, V.V. Calmidi, R.L. Mahajan, Thermophysical properties of high
porosity metal foams, Int. J. Heat Mass Transfer 45 (2002) 10171031.
[19] A. Bhattacharya, R.L. Mahajan, Finned metal foam heat sinks for electronics
cooling in forced convection, Int. J. Electron. Packag. 124 (2002) 155163.
[20] K. Boomsma, D. Poulikakos, F. Zwick, Metal foams as compact high
performance heat exchangers, Mech. Mater. 35 (2003) 11611176.
[21] I. Kurbas, N. Celik, Experimental investigation on forced and mixed convection
heat transfer in a foam-lled horizontal rectangular channel, Int. J. Heat Mass
Transfer 52 (2009) 13131325.

884

N. Dukhan et al. / International Journal of Heat and Mass Transfer 67 (2013) 877884

[22] S.Y. Kim, B.H. Kang, J.-H. Kim, Forced convection from aluminum foam
materials in asymmetrically heated channel, Int. J. Heat Mass Transfer 44
(2001) 14511454.
[23] C.Y. Zhao, T. Kim, T.J. Lu, H.P. Hodson, Thermal transport in high porosity metal
cellular metal foams, J. Themophys. Heat Transfer 18 (3) (2004) 309317.
[24] S.-C. Tzeng, T.-M. Jeng, Convective heat transfer in porous channels with 90deg turned ow, Int. J. Heat Mass Transfer 49 (2006) 14521461.
[25] G. Hetsroni, M. Gurevich, R. Rozenblit, Metal foam heat sink for transmission
window, Int. J. Heat Mass Transfer 48 (2005) 37933803.
[26] J.-S. Noh, K.B. Lee, C.G. Lee, Pressure loss and forced convective heat transfer in
an annulus lled with aluminum foam, Int. J. Heat Mass Transfer 33 (2006)
434444.
[27] S.Y. Kim, J.W. Paek, B.H. Kang, Flow and heat transfer correlations for porous
n in a plate-n heat exchanger, J. Heat Transfer 122 (2000) 572578.
[28] N. Dukhan, Analysis of Brinkman-extended Darcy ow in porous media and
experimental verication using metal foam, J. Fluids Eng. 134 (7) (2012),
http://dx.doi.org/10.1115/1.4005678.
[29] S. Krishnan, J.Y. Murthy, S.V. Garimella, A two temperature model for the
analysis of passive thermal control system, J. Heat Transfer 126 (2004) 628
637.
[30] A. Amiri, K. Vafai, Analysis of dispersion effects and non-thermal equilibrium,
non-Darcian, variable porosity incompressible ow through porous media, Int.
J. Heat Mass Transfer 37 (6) (1994) 939954.
[31] G.J. Hwang, C.H. Chao, Heat transfer measurement and analysis for sintered
porous channels, J. Heat Transfer 116 (1994) 456464.
[32] A. Amiri, K. Vafai, T.M. Kuzay, Effects of boundary conditions on non-Darcian
heat transfer through porous media and experimental comparisons, J. Numer.
Heat Transfer Part A 27 (1995) 651664.
[33] B. Alazmi, K. Vafai, Constant wall heat ux boundary conditions in porous
media under local thermal non-equilibrium conditions, Int. J. Heat Mass
Transfer 45 (2002) 30713087.

[34] S.J. Kim, D. Kim, The thermal interaction at the interface between a porous
medium and an impermeable wall, J. Heat Transfer 123 (2001) 527533.
[35] G.E. Myers, Analytical Methods in Conduction Heat Transfer, second ed.,
AMCHT Publications, Madison, WI, 1998. p. 46.
[36] V.S. Arpaci, Conduction Heat Transfer, Addison-Wesley, Reading, MA, 1966. pp.
135136.
[37] N. Dukhan, K. Hooman, Comments on two analyses of thermal nonequilibrium DarcyBrinkman convection in cylindrical porous media, Int. J.
Heat Mass Transfer 66 (2013) 440443.
[38] ERG Materials and Aerospace: http://www.ergaerospace.com, accessed March,
2013.
[39] R. Figliola, D. Beasley, Theory and Design for Mechanical Measurements, John
Wiley and Sons, New York, NY, 2000. pp. 149163.
[40] F.M. White, Fluid Mechanics, fourth ed., McGraw-Hill, New York, 1999.
[41] V.V. Calmidi, R.L. Mahajan, The effective thermal conductivity of high porosity
brous metal foams, J. Heat Transfer 121 (1999) 466471.
[42] N. Dukhan, M. Ali, Strong wall and transverse size effects on pressure drop of
ow through open-cell metal foam, Int. J. Therm. Sci. 57 (2012) 8591.
[43] N. Dukhan, P. Patel, Equivalent particle diameter and length scale for pressure
drop in porous metals, Exp. Therm. Fluid Sci. 32 (2008) 10591067.
[44] F. Kuwahara, M. Shirota, A. Nakayama, A numerical study of interfacial
convective heat transfer coefcient in two-energy equation model for
convection in porous media, Int. J. Heat Mass Transfer 44 (2001) 11531159.
[45] M. Sahraoui, M. Kaviany, Slip and no-slip temperature boundary condition at
the interface of porous/plain media, Int. J. Heat Mass Transfer 36 (4) (1993)
10191033.
[46] K. Vafai, K. Yang, A Note on the thermal non-equilibrium in porous media and
heat ux bifurcation phenomenon in porous media, Transp. Porous Med. 96
(2013) 169172.

S-ar putea să vă placă și