You are on page 1of 9

Applied Thermal Engineering 37 (2012) 1e9

Contents lists available at SciVerse ScienceDirect

Applied Thermal Engineering


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

Experimental and numerical studies on melting phase change heat transfer


in open-cell metallic foams filled with paraffin
W.Q. Li, Z.G. Qu*, Y.L. He, W.Q. Tao
MOE Key Laboratory of Thermo-Fluid Science and Engineering, Xi’an Jiaotong University, 28# XianNing Road, 710049 Xi’an, China

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

Article history: In the current study, the melting phase change heat transfer in paraffin-saturated in open-celled metallic
Received 28 July 2011 foams was experimentally and numerically studied. The experiments were conducted with seven high-
Accepted 2 November 2011 porosity copper metal foam samples (3  90%), and paraffin was applied as the phase-change material
Available online 17 November 2011
(PCM). The wall and inner temperature distribution inside the foam were measured during the melting
process. The effects of foam morphology parameters, including porosity and pore density, on the wall
Keywords:
temperature and the temperature uniformity inside the foam were investigated. The melting heat
Phase-change material
transfer is enhanced by the high thermal conductivity foam matrix, although its existence suppresses the
Metallic foam
Melting
local natural convection. A numerical model considering the non-Darcy effect, local natural convection,
Non-equilibrium model and thermal non-equilibrium was proposed. The velocity, temperature field, and evolution of the solid
eliquid interface location at various times were predicted. The numerically predicted results are in good
agreement with the experimental findings. The model as well as the feasibility and necessity of the
applied two-equation model were further validated.
Ó 2011 Elsevier Ltd. All rights reserved.

1. Introduction electronic devices. For example, in some electronic cooling systems,


the chip experiences a transient or periodic heat generation, which
Phase-change materials (PCMs) have been widely used in many requires a highly efficient coolant system to dissipate the heat in
applications, such as passive cooling for electronic devices, protec- case of chip exposure to extreme-temperature environments.
tion systems in aircrafts, food processing, and energy conservation in However, the poor conductivity of organic PCMs reduces the rate of
buildings, because of their high latent heat, chemical stability, suit- heat storage, thereby increasing the junction temperature of the
able phase-change temperature, and reasonable price. Experimental devices beyond the allowable range.
and analytical/numerical studies in published literatures have In order to address this unacceptable problem, thermal
focused on moving boundary problems [1e4]. Agyenim et al. [5] conductivity enhancement techniques that increase heat transfer
summarized the various applications of PCMs with different rates have been developed. These enhancement techniques are
melting temperatures in suitable thermal energy storage systems. summarized in the following three methods: (1) Dispersing high-
The size and the shape of the PCM container were also taken into conductivity particles in PCMs; Wang et al. [7] experimentally
consideration to ensure the long-term stable thermal performance of proved that the thermal conductivity of composite PCMs is
the system. Dutil et al. [6] presented main four types of numerical enhanced by incorporating a b-aluminum nitride additive. (2)
solutions dealing with the thermal behaviors of PCM in solid/liquid Utilizing high-conductivity matrices, such as a metal or a graphite
systems and showed the predicted results of different configurations. compound, as heat delivery promoters; Kim and Drazal [8]
Although some organic PCMs, such as paraffin, are very popular improved the effective thermal conductivity of paraffin by stirring
in energy storage applications and electronic cooling systems exfoliated graphite nanoplatelets (xGnP) in liquid paraffin. The
because of the aforementioned advantages and their low density authors found that the thermal conductivity of paraffin/xGnP
compared with other kinds of heat storage materials (e.g., metal composite PCMs increases as the xGnP loading content increases
PCMs and hydrated salts). However, organic PCMs suffer from low without reducing the latent heat of the paraffin wax. (3) Filling
conductivity (z0.1 W m1 K1), which is likely to cause failure in extensive surfaces such as fins into the body of PCM; Shatikian et al.
[9] numerically investigated the effect of internal fins on melting
* Corresponding author. Tel./fax: þ86 029 82668036.
rate, melting front profiles, and heat transfer. (4) Varying the shape
E-mail address: zgqu@mail.xjtu.edu.cn (Z.G. Qu). of the heat storage-heat transfer system; Banaszek et al. [10].

1359-4311/$ e see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.applthermaleng.2011.11.001
2 W.Q. Li et al. / Applied Thermal Engineering 37 (2012) 1e9

Nomenclature T temperature ( C)
Tm1 the lower limit of melting point
A additional term in the momentum equation Tm2 the upper limit of melting point
asf interfacial surface area (m1) t time (s)
C constant parameter u,v velocity in x and y directions (m s)
CI inertial coefficient x,y cartesian coordinates
cpf specific heat capacity of PCM (J kg1 K1)
cps specific heat capacity of metal foam (J kg1 K1) Greek symbols
ctd thermal dispersion coefficient r density (kg m3)
df fiber diameter (m) d liquid fraction (¼3 fl)
dp pore size (m) b thermal expansion coefficient (K1)
dk characteristic length (m) m dynamic viscosity (N s m2)
fl liquid fraction in the pore 3 porosity
g gravitational acceleration (m s2) c tortuosity coefficient
hsf interfacial heat transfer coefficient (W m2 K1) u pore density (pore per inch, PPI)
K permeability (m2)
k thermal conductivity (W m1 K1) Subscripts
ke effective thermal conductivity (W m1 K1) amb ambient
kf thermal conductivity of fluid (W m1 K1) f fluid (both solid and liquid paraffin)
kfe effective thermal conductivity of fluid (W m1 K1) fl liquid paraffin
ks thermal conductivity of solid (W m1 K1) fs solid paraffin
kse effective thermal conductivity of metal foam m melting
(W m1 K1) p pore
L latent heat of paraffin (J kg1 K1) s solid matrix
Nu Nusselt number sf surface
P pressure (Pa) td thermal dispersion
PPI pore number per inch
Pr Prandtl number Superscripts
q heat flux (W m2) n iteration number at the present time level
Re Reynolds number

designed an energy conservation system using helix-shaped influences of the Rayleigh, Stefan, and Nusselt numbers on the
channels where PCM and heat transfer fluids were separately evolution of the solid/liquid interface were reported and discussed.
enclosed in every other channel. Such a design makes it possible for Although this technique can solve more complicated cases, the
the fluid to stick to the channel wall more strictly when flowing thermal dispersion effects were not considered, thereby making
through the channel because of centrifugal force and more effi- the technique unsuitable for cases with a high Rayleigh number.
ciently take away heat from the PCM. However, a disadvantage of Moreover, the effective thermal conductivity is only determined by
such a structure for heat exchange is the increased consumption the volume fraction of each phase without considering the
and complexity of the cooling system. geometric configuration effect of the metal structure, which is not
In addition to the aforementioned methods, inserting PCMs into suitable for metal foam. Lafdi et al. experimentally investigated the
a porous media appears to be an attractive choice to enhance heat phase-change heat transfer within a PCM composite permeated
transfer. A number of experimental and theoretical investigations with aluminum foam [15], and then numerically studied the cool-
on this problem have been conducted in the past two decades. ing process of the electronic device using heat sinks of solid matrix
Beckermann and Viskanta [11] conducted both experimental and impregnated with PCM [16]. Zhao et al. [17] experimentally and
numerical studies on the melting of gallium in glass beads. In their numerically studied the paraffin melting in metallic foam.
numerical study, the one-equation energy model was adopted to However, in their model, the flow motion of liquid paraffin, which
predict the temperature, taking into account the natural convective has been proved to be very important, was not taken into account.
effect. To the authors’ best knowledge, the one-equation model is Jegadheeswaran and Plhekar [18] have summarized various tech-
suitable only when the thermal properties of the solid matrix and niques for enhancing the thermal performance in different heat
the saturated PCM are in the same order of magnitude; that is, the thermal storage systems and enumerated the pros and cons of
macroscopic temperatures of the two materials should be close these enhancement techniques.
enough so that a single temperature model can describe the whole The aim of the present work was to experimentally and
melting process. This equilibrium model fails when the thermal numerically investigate the melting process of paraffin in copper
properties of the two materials are significantly different or when foams. Metal foam, a metallic porous matrix material, is a porous
the convective effect becomes vitally important [12]. Extensive media that exhibits the excellent combination of compactness, low
studies are conducted to improve the accuracy of the model. weight, and high thermal conductivity [19]. An experimental test
Harris et al. [13] theoretically studied the phase-change process in rig was built to measure the wall and internal temperatures. The
porous media using a two-temperature model. However, the main effects of porosity and pore density on heat transfer were examined
drawback of this approach is its difficulty in handling complex and discussed. A two-equation model was subsequently applied to
cases, such as buoyancy-driven natural convection in a molten numerically study the solideliquid phase-change process in high-
region in porous media. Krishnan et al. [14] numerically investi- porosity metallic foam. In the model, the natural convection of
gated the thermal transport phenomena associated with phase liquid paraffin, the thermal dispersion effect, the irregular geom-
change in a rectangular cavity filled with metal foam. The etry configuration, and the non-Darcy effects were fully considered.
W.Q. Li et al. / Applied Thermal Engineering 37 (2012) 1e9 3

The numerical predictions were validated by comparing the Table 1


experimental results. The velocity field of the molten paraffin, the Thermal physical properties of paraffin.

temperature distributions for PCM and the metal matrix, as well as Parameters Value
the time-related interface variations were presented and discussed. rf (at 20  C) 785.02
L 1.021  105
Tm1 46.48
2. Experimental apparatus and procedure Tm2 60.39
kfs 0.30
2.1. PCM melting point and latent heat measurement kfl 0.10
cpf (at 20  C) 2850
m (at 65  C) 3.65  103
The latent heat and melting point of the paraffin were measured b (at 65  C) 3.085  104
using a differential scanning calorimeter (DSC; TA-Q20). The results
shown in Fig. 1 indicate that the latent heat of paraffin is 102.1 J/g,
and the melting process starts at 46.48  C and ends at 60.39  C. The inside the hole at 3, 13, 28, and 43 mm from the left side of the
related thermophysical properties were listed in Table 1. metal foam. The uncertainties were estimated based on the random
errors during temperature measurements. The uncertainty for the
T-type thermal couple temperature is 0.1  C. The uncertainties for
2.2. Experimental apparatus and procedure
the D.C. power output and the film heater power output are 1.22%
and 5.0%, respectively.
As shown in Fig. 2, a plexiglass cavity was designed to cage the
The morphology of the metal foam used in the experiment is
paraffin-saturated metal foam based on the size of the metal foam.
shown in Fig. 3. Seven copper foam samples with different poros-
However, considering the volume expansion of PCM during phase
ities and pore densities were used, the geometric characteristics of
change, a 5 mm gap between the top surface and the inner glass
which are listed in Table 2.
surface was left for the expansion of the melted paraffin. In case of
leakage during melting, two other 30 mm thick plexiglass plates
were fabricated and fixed to the left and right sides of the cavity and 3. Experimental results and discussion
tightened at the corners with four studs. All copper foam samples
having a uniform dimension of 100 mm  100 mm  45 mm were 3.1. Effect of the metal matrix and foam porosity
sintered on a 2 mm thick copper substrate at the bottom to improve
the thermal conduction to the internal space and minimize the Fig. 4 displays the wall temperature variation history with time
contact thermal resistance. A 100 mm  100 mm  0.15 mm of three metal foams with different porosities (3 ¼ 0.90, 0.95, and
electric insulated film heater was glued to the left side of the metal 0.98), with the pore density fixed at 40 PPI. The wall temperature
foam, i.e., a brazed copper substrate, to provide constant heat flux. variation of pure PCM was used as the reference. The whole melting
The other four sides (top, bottom, front, and back) were kept process can be divided into three regions: solid, mush, and pure
thermally insulated with 30 mm thick urethane foam plates. A liquid. In the solid region, the temperatures for pure PCM and foam-
constant heat flux of 4000 W/m2 was supplied by a DC voltage PCM composite both increased by absorbing explicit heat from the
stabilizer to the film heater during the experiment. For each wall. The foam-PCM composite wall temperature was lower than
sample, a temperature-tracking task was completed by a Keithley that of pure PCM and showed better thermal management ability
2700-multimeter/data acquisition system and data were obtained because the effective PCM thermal conductivity was improved by
every 5 min. A small cylindrical hole (3 mm in diameter, 45 mm in the copper metal matrix. As time progressed, the metal ligaments
height) erected on the sintered copper plate was drilled in the reached a temperature higher than the PCM melting point and
center of the foam to better understand the time-related variation started triggering a local phase change. Two factors dominate the
in the inner temperatures. Five T-type thermocouples were phase change, namely, the heat conduction and the natural
selected, as seen in Fig. 2(b). One was attached to the copper plate convection of the molten PCM. The final performance depends on
to measure the wall temperature, and the other four were fixed whichever of these two factors prevails in the melting process. For
the pure PCM, the paraffin absorbed the latent heat at the solid/
liquid melting interface dominated by natural convection, thereby
increasing the temperature of pure paraffin in a flattened manner.
However, in the foam-PCM composite, the heat conduction thermal
resistance was evidently reduced by the foam matrix, although the
liquid paraffin was constrained in the metal foam pore to some
extent and suppressed natural convection, which plays a positive
role in heat transfer by accelerating the melting rate. Consequently,
heat conduction dominated the melting process, and the wall
temperature of the foam-PCM composite was lowered and
increased almost linearly, governed by the combined effects of heat
diffusion, natural convection, and the latent heat absorption at the
moving boundary. When the melting process was fully complete,
the explicit heat exerted its role again and the temperatures for
both pure PCM and the composite continued to increase linearly.
Points A, B, C, and D represented the end of the melting process in
the mush region of the composites with 0.90, 0.95, and 0.98
porosities and pure PCM (3 ¼ 1). The total time consumption
decreased in the order of decreasing porosity because a higher
porosity indicates a higher PCM volume fraction. Higher porosity
Fig. 1. DSC results for paraffin. means a lower effective thermal conductivity at a fixed pore
4 W.Q. Li et al. / Applied Thermal Engineering 37 (2012) 1e9

Fig. 2. Experimental apparatus and test section. (a) Experimental apparatus, (b) Test section.

density, and implies a smaller fiber diameter or smaller heat increasing pore density at the same time; however, the influence
transfer surface area for the prevailing heat conduction process. was less significant for the pore density compared with the
Thus, higher porosity allows the sample to achieve higher wall porosity. As pore density increased, the interfacial surface area and
temperature. The effective thermal conductivity can be calculated the effective thermal conductivity increased, thereby improving
according to Ref. [20]. The interfacial surface area can be obtained the melting heat transfer; nevertheless, the permeability
from Eq. (1) [21]: decreased, resulting in a strong suppression of the natural
convection of molten wax. The two opposite factors competed and
3pdf strong suppression of the natural convection is comparatively
asf ¼ (1)
d2p considerable, resulting in higher wall temperature for high pore
density foam and a less evident temperature difference among the
three composites.
3.2. Effect of pore density

Fig. 5 depicts the relationship between the wall temperature


and the duration time for the copper foam filled with PCM of three Table 2
pore densities (10, 20, and 40 PPI) and pure PCM with a fixed 0.9 Geometrical parameters of the sample foams.
porosity. A similar trend was observed when the wall temperature Sample Porosity Pore density (PPI) dp (mm) df (mm)
for the copper foam filled with PCM was lower than that of pure 1 0.90 10 2.540 0.309
PCM. The total melting duration time at various pore densities were 2 0.90 20 1.270 0.154
almost the same because the PCM volume fraction were identical at 3 0.90 40 0.635 0.077
a fixed porosity. The wall junction temperature became higher with 4 0.95 10 2.540 0.218
5 0.95 20 1.270 0.109
6 0.95 40 0.635 0.055
7 0.98 40 0.635 0.035

Fig. 4. Variations of Tw with time of PCM composites with different porosities at a pore
Fig. 3. Typical structures of tested metal foam (3 ¼ 0.90, u ¼ 40 PPI). density of 40 PPI.
W.Q. Li et al. / Applied Thermal Engineering 37 (2012) 1e9 5

PPI foam than for the 40 PPI foam. The motion of molten paraffin wax
in the 10 PPI foam was less restricted in the matrix because of the
larger pore size, leading to a stronger natural convective effect.
Consequently, the temperature distribution was more uniform in the
horizontal direction. The effect of porosity on temperature distribu-
tion uniformity is related to pore density because the total thermal
resistance is determined by both the effective thermal conductivity
and natural convection. The natural convection was severely sup-
pressed at the 40 PPI pore density, whereas it was considerable at the
10 PPI pore density. When the porosity varied from 0.95 to 0.9, the
effective thermal conductivity increased or the heat conduction
thermal resistance decreased. The influence of the effective thermal
conductivity increment by decreasing the porosity from 0.95 to 0.9 on
the total resistance was evident for the 40 PPI foam compared with
the 10 PPI foam. Hence, the improvement in temperature distribution
uniformity is more significant for the 40 PPI foam than for the 10 PPI
foam, as shown in Fig. 6(a)e(d).

Fig. 5. Variations of Tw with time for the PCM composites with different pore densities 4. Numerical simulation
(3 ¼ 0.90).

4.1. Physical and mathematical model


3.3. Internal temperature distributions
The two-dimensional schematic diagram of the present tran-
Fig. 6 quantitatively presents the internal temperature distribution
sient melting heat transfer is shown in Fig. 7. The copper foam
at four different foam composite test spots at two porosities (3 ¼ 0.90
saturated with solid paraffin measured 45 mm in length and
and 0.95) and pore densities (u ¼ 10 and 40 PPI). A comparison of
100 mm in height. The top, bottom, and right sides can be
Fig. 6(a) and (b) or Fig. 6(c) and (d) shows that the temperature
considered adiabatic. The initial temperature was set as the
difference between the two adjacent tested spots is smaller for the 10
ambient temperature at 25  C. The left side was heated at

a b

c d

Fig. 6. Internal temperature-time variation.


6 W.Q. Li et al. / Applied Thermal Engineering 37 (2012) 1e9

Continuous equation:

vrf  !
þ V$ rf h U i ¼ 0 (2)
vt
Momentum equations:
 
rf vhui rf ! vP m m rf CI !
þ 2 ðh U i$VÞhui ¼  þ fl V2 hui  fl þ pffiffiffiffi jh U ij hui
d vt d vx d K K
þ Ahui ð3Þ

 
rf vhvi rf ! vP mfl 2 mfl rf CI !
þ 2 ðh U i$VÞhvi ¼  þ V hvi  þ pffiffiffiffi jh U ij hvi
d vt d vy d K K
 
þ rf g b Tf  Tm1 þ Ahvi ð4Þ

Energy equation for PCM:

! D E
Fig. 7. Schematic diagram of the physical domain.
dfl v Tf ! D E
3 rf cpf þ L D E þ rf cpf ðh U i$VÞ Tf
a constant heat flux of 4000 W m2, consistent with the input d Tf vt
power of the heating film. The following assumptions are made in   D E  D E
the mathematical model for the present problem: (1) the metal ¼ kfe þ ktd V2 Tf þ hsf asf hTs i  Tf (5)
foam is assumed homogeneous and isotropic; (2) the liquid PCM is
Energy equation for metal matrix:
considered impressible and Newtonian, and subjected to the
Boussinesq approximation; (3) the changes of density and heat vhTs i  D E
capacity of paraffin during the whole melting process is neglected ð1  3 Þrs cps ¼ kse V2 hTs i  hsf asf hTs i  Tf (6)
vt
while the other thermal physical properties of PCM are constant at
each phase but treated separately for the solid and liquid phases. Where thermal properties ofrf, L,cpf,mfl,b are presented in Table 1.
Based on the above-mentioned considerations, the volume- The properties ofrs (8920 kg m-3), cps (380 J kg-1 K-1) and ks
averaging continuous, momentum, and two-energy equations for (401 W m-1 K-1)are density,heat capacity and the thermal
PCM and the solid matrix can be written as follows: conductivity of copper, respectively. Other parameters in the

Table 3
The employed semi-empirical correlations of metallic foam parameters.

Parameter Correlation Reference


Permeability, K 3
2 d2 [23]
k
K ¼
36ðc  1Þc
 
4p 1
c ¼ 2 þ 2cos þ cos1 ð23  1Þ
3 3
c 22:4  103
dk ¼ dp dp ¼
3c u
pffiffiffiffi
Inertial coefficient, CI FI ¼ CI ð1  3 Þ0:132 ðdf =dp Þ1:63 = K , CI ¼ 2.12  103 [26]
rffiffiffiffiffiffiffiffiffiffiffi
13
df ¼ 1:18 dp
3p
( 0:76Re0:4 Pr0:37 k =d ; 1  Re  40
d f f d
hsf ¼ 0:52Re0:5 d Pr
0:37
kf =df ; 40  Red  103
0:26Re0:6 d Pr
0:37
kf =df ; 103  Red  2  105
Local heat transfer coefficient, hsf [24]
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Red ¼ rf u2 þ v2 df =ð3 mfl Þ
Thermal dispersion conductivity, ktd Ctd pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi [25]
ktd ¼ r C d u2 þ v2 , Ctd ¼ 0.36
1  3 f pf f
1
ke ¼ pffiffiffi
2ðRA þ RB þ RC þ RD Þ
4l
RA ¼
ð2e2 þ plð1  eÞÞks þ ð4  2e2  plð1  eÞÞkf
ðe  2lÞ2
RB ¼
ðe  2lÞe2 ks þ ð2e  4l  ðe  2lÞe2 Þkf
pffiffiffi
ð 2  2eÞ2
RC ¼ pffiffiffi pffiffiffi pffiffiffi
2pl ð1  2 2eÞks þ 2ð 2  2e  pl ð1  2 2eÞÞkf
2 2

Effective thermal conductivity, ke 2e [20]


RD ¼ 2
e ks þ ð4  e2 Þkf
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffi pffiffiffi
2ð2  ð5=8Þe3 2  23 Þ
l ¼ pffiffiffi , e ¼ 0.339
pð3  4 2e  eÞ
( kfs;e solid region
kfe ¼ fl kfl;e þ ð1  fl Þkfs;e Mush region
kfl;e liquid region
kse ¼ ke jkf ¼0 , kfe ¼ ke jks ¼0
W.Q. Li et al. / Applied Thermal Engineering 37 (2012) 1e9 7

governing equations, including K (permeability), CI (inertial coef- Table 4


ficient), kse (the effective thermal conductivity of the solid matrix), Initial and boundary conditions for the governing equations.

kfe (the effective thermal conductivity of paraffin), ktd (thermal Definite conditions Positions Velocity Temperature
dispersion conductivity), and hsf (local heat transfer coefficient), are Initial condition 0  x  L, 0  y  H u¼v¼0 Tf ¼ Ts ¼ Tamb
from the semi-empirical correlations in the literature and can be x ¼ 0, 0  y  H u¼v¼0 q ¼ qfþ qs
seen in Table 3, where the values of CI and Ctd shown the best fit to ¼3 q þ (1  3 )q
x ¼ L, 0  y  H u¼v¼0 vTf vTs
the experimental data in the corresponding literatures. For inter- ¼ ¼ 0
vx vx
facial heat transfer coefficienthsf, since no exact correlation is Boundary conditions 0  x  L, y ¼ 0 u¼v¼0 vTf vTs
¼ ¼ 0
suitable for natural convection in porous media, an empirical vy vy
correlation provided by Zukauska [24] has been introduced. The 0  x  L, y ¼ H u¼v¼0 vTf vTs
¼ ¼ 0
Reynolds number is based on the cylinder diameter and Pr is the vy vy

fluid Prandtl number. In this study, the Re range was verified to be


less than 5.0 which satisfied the first term of correlation propased
by Zukauska [24] and then was applied. The bracket “hi” represents 5. Numerical results and discussion
the volume-averaging model based on the DupuiteForchheimer
relationship [22]. The second and fourth terms on the right side of 5.1. Code validation
Eqs. (3) and (4) account for the extension of Darcy’s law to explain
the non-Darcy effects, and the boundary and inertial effects The numerical codes were validated with the present experi-
proposed by Brinkman and Forchheimer, respectively. The last mental results by comparing the interface positions at three
terms in the momentum equations are additional source terms, different moments. Fig. 8(a) and (b) are pictures of interface loca-
where parameter A is related to the liquid fraction in the pore tions captured by a digital camera at 3600 s and 3780 s for the 0.90
volume based on Kozeny’s equation [22], which is introduced as the porosity and 10 PPI foam. The predicted solid/liquid interfaces
following equation: [Fig. 8(c) and (d)] obtained from the numerical predictions agree

 
C 1  fl2
A ¼ (7)
S þ fl3

8 0 Tf  Tm1
< .
fl ¼ Tf  Tm1 ðTm2  Tm1 Þ Tm1  Tf  Tm2 (8)
:
1 Tm2  Tf

d ¼ 3 $fl (9)

where d is the liquid fraction considered as a function of PCM


temperature and porosity, and C and S are set as 1015 and 1010,
respectively, to fix the PCM velocity to zero before melting for
convenient numerical implementation. Therefore, the conduction
dominated equations for the solid phase and convection-diffusion
controlled equations for melted paraffin are unified by Eqs.
(3)e(5). The initial and boundary conditions for the governing
equations are given in Table 4. The initial temperatures for PCM and
metal foam are equal to the ambient temperature. The initial values
of velocity for PCM at both directions are zero and the same value is
applied to the velocity boundary condition based on the velocity
non-slip principle at the wall. At the left boundary, the cover plate
contacts with metal fiber or the fluid (solid or liquid), due to this
consideration, if ignoring the heat dissipation through the sintered
2 mm copper substrate, the heat is assumed distributed respec-
tively to the metal foam and PCM by each representative volume
ratio at the wall surface as applied in [26]. The rest boundaries are
set adiabatic.

4.2. Numerical procedure

The governing equations were discretized by a finite volume


method in a staggered grid system of 90  80. The combined
equations are solved numerically using the IDEAL algorithm [27].
The regions of solid paraffin and melted region are determined
automatically during the iteration by the whole domain solution
strategy. The iteration at the present time layer is considered
convergent if the maximum relative residuals of Tf are less than Fig. 8. Comparison of the solid/liquid interface between the photos and numerical
105, which can be expressed as ½ðTfn  Tfn1 Þ=Tfn1  < 105 . results. (3 ¼ 0.90,u ¼ 10 PPI) at different times.
8 W.Q. Li et al. / Applied Thermal Engineering 37 (2012) 1e9

Fig. 9. Velocity field and temperature distributions (3 ¼ 0.95, u ¼ 20 PPI) at 2700 s.

well with the experimental results. The feasibility of the present whereas only diffusion existed in the solid paraffin region. A
model is verified. difference in the temperature distributions between PCM and the
metal matrix was observed, especially at the liquid paraffin region.
5.2. Temperature distribution This quantitative result validates the feasibility and necessity of the
two-temperature model.
Fig. 9 shows the velocity field and temperature contours of PCM
and the solid matrix (3 ¼ 0.95, u ¼ 20 PPI) at 2700 s. The hotter 5.3. Interface variations
liquid moved upward because of the natural convection of molten
paraffin stimulated by buoyancy, pushing the upper portion of the Fig. 10(a) and (b) compare the time-related interface position for
interface faster than the bottom. For the temperature distribution, the 0.9 and 0.95 porosities at a fixed 20 PPI pore density. At the
it could be clearly observed that there exist great differences beginning of the melting process, the interface tended to be vertical
between metal foam and the PCM. Also, at the same local position, because the liquid fraction was relatively small, indicating that heat
metal foam temperature was higher than that of PCM. In the solid conduction dominated in the process. As melting continued, the
PCM region, conduction-induced heat transfer produced similar upper part moved faster than the lower part because of the eddy
temperature profiles for the PCM and the metallic matrix; however, flow, promoting the fusion rate at the top and the phase change was
the metallic foam temperature was higher than that of the solid accelerated because of the natural convection induced by buoyancy
PCM due to the local thermal non-equilibrium effect. In the liquid in the molten PCM. Moreover, the interface movement rate for the
wax region, the liquid temperature pattern differs from that of the 0.90 porosity foam was higher than that of the 0.95 porosity foam.
metallic matrix. The temperature profile of the liquid paraffin was At 1500 s, the top phase interface for the 0.90 porosity foam had
more uniform in the horizontal direction than that of the metallic already attached to the right boundary, whereas the interface of the
matrix because natural convection prevailed in the liquid region, 0.95 porosity foam had just crossed over the middle (X/H ¼ 0.57).

a b 0.10
0.10

0.08 0.08

0.06 0.06
600s
Y
Y

900s
0.04 600s 0.04
1200s
900s 1500s
1200s 1800s
0.02 1500s 0.02 2100s
1800s 2400s
2100s 3000s
0.00 0.00
0.005 0.010 0.015 0.020 0.025 0.030 0.035 0.040 0.045 0.005 0.010 0.015 0.020 0.025 0.030 0.035 0.040 0.045
X X
ε=0.90 ε=0.95
Fig. 10. Solid/liquid interface locations at various times (u ¼ 20 PPI).
W.Q. Li et al. / Applied Thermal Engineering 37 (2012) 1e9 9

The above-mentioned results are consistent with the experimental [4] P. Lamberg, R. Lehtiniemi, A.M. Henell, Numerical and experimental investi-
gation of melting and freezing processes in phase change material storage, Int.
results indicated in Fig. 4.
J. Therm. Sci. 43 (2004) 277e287.
[5] F. Agyenima, N. Hewitt, P. Eames, M. Smyth, A review of materials, heat transfer
and phase change problem formulation for latent heat thermal energy storage
6. Conclusion systems (LHTESS), Renew. Sust. Energ. Rev. 14 (2010) 615e628.
[6] Y. Dutil, D.R. Rousse, N.B. Salah, S. Lassue, L. Zalewski, A review on phase-
The melting phase-change heat transfer in paraffin PCM was change materials: mathematical modeling and simulations, Renew. Sust.
Energ. Rev. 15 (2011) 112e130.
enhanced by the porous metallic foam. The mechanisms of heat [7] W. Wang, X. Yang, Y. Fang, J. Ding, J. Yan, Enhanced thermal conductivity and
diffusion and natural convection dominated the melting phase- thermal performance of form-stable composite phase change materials by
change heat transfer. The enhancement of heat conduction is using b-Aluminium nitride, Appl. Energy 86 (2009) 1196e1200.
[8] S. Kim, L.T. Drazal, High latent heat storage and high thermal conductive
more prominent than the suppression of natural convection of phase change materials using exfoliated graphite nanoplatelets, Sol. Energy
molten liquid by the copper matrix in the foam-PCM composite, Mater. Sol. Cells 93 (2009) 136e142.
whereas natural convection prevailed in the pure PCM. The final [9] V. Shatikian, G. Ziskind, R. Letan, Numerical investigation of a PCM-based heat
sink with internal fins, Int. J. Heat Mass Transfer 48 (2005) 3689e3706.
total thermal resistance was lower for the foam-PCM composite [10] J. Banaszek, R. Domañski, M. Rebow, F. El-Sagier, Numerical analysis of the
than the pure PCM to result in the lower corresponding wall paraffin waxeair spiral thermal energy storage unit, Appl. Therm. Eng. 19
temperature. The influence of pore density on wall temperature (1999) 1253e1277.
[11] C. Beckermann, R. Viskanta, Natural convection solid/liquid phase change in
was less sensitive than that of porosity. The uniformity of the porous media, Int. J. Heat Mass Transfer 31 (1988) 35e46.
temperature distribution inside the foam-PCM composite was [12] M. Kaviany, Principles of Heat Transfer in Porous Media (Mechanical Engi-
augmented either by decreasing pore density to accelerate natural neering Series). Springer, Berlin, 1995.
[13] K.T. Harris, A. Haji-Sheikh, A.G. Agwu Nnanna, Phase-change phenomena in
convection or by decreasing porosity to improve the effective
porous media-a non-local thermal equilibrium model, Int. J. Heat Mass
thermal conductivity. The feasibility of the numerical model, in Transfer 44 (2001) 1619e1625.
which the non-Darcy effects, the natural convection in the liquid [14] S. Krishnan, J.Y. Murthy, S.V. Garimella, A two-temperature model for solid-
paraffin, and the local non-equilibrium effects were considered, liquid phase change in metal foams, J. Heat Transfer 127 (2005) 995e1004.
[15] K. Lafdi, O. Mesalhy, S. Shaikh, Experimental study on the influence of foam
was validated by the experimental visualization. The temperature porosity and pore size on the melting of phase change materials, J. Appl. Phy.
distribution profile for the local solid matrix was quite different 102 (2007) 083549.
from that of the filled PCM and validated the feasibility and [16] K. Lafdi, O. Mesalhy, A. Elgafy, Merits of employing foam encapsulated phase
change materials for pulsed power electronics cooling applications, J. Elec.
necessity of the applied two-equation model. The solid/liquid Packing 130 (2008) 021004.
movement rate was higher for the low porosity foam, which is also [17] C.Y. Zhao, W. Lu, Y. Tian, Heat transfer enhancement for thermal energy
consistent with the experimental results. storage using metal foams embedded within phase change materials (PCMs),
Sol. Energy 84 (2010) 1402e1412.
[18] S. Jegadheeswaran, S.D. Pohekar, Performance enhancement in latent heat
thermal storage system: a review, Renew. Sust. Energ. Rev. 13 (2009)
Acknowledgements 2225e2244.
[19] M.F. Ashby, T.J. Lu, Metal foams: a survey, Sci. China (Series B.) 46 (2003) 521e532.
The current study was supported by the National Key Projects of [20] K. Boomsma, D. Poulikakos, On the effective thermal conductivity of a three-
dimensionally structured fluid-saturated metal foam, Int. J. Heat Mass
Fundamental R/D of China (973 Project: 2011CB610306), the
Transfer 44 (2001) 827e836.
National Natural Science Foundation of China (Nos. 51176149 and [21] V.V. Calmidi, R.L. Mahajan, Forced convection in high porosity metal foams,
50736005), the National Excellent Doctoral Dissertation Founda- J. Heat Transfer 122 (2000) 557e565.
[22] A. Donald Nield, Convection in Porous Media. Springer, New York, 2006.
tion of China (201041), the Doctoral Fund of the Ministry of
[23] J.G. Fouriea, J.P. Du Plessis, Pressure drop modeling in cellular metallic foams,
Education of China (200806981013). Chem. Eng. Sci. 57 (2002) 2781e2789.
[24] A. Zhukauskas, Heat Transfer from Tubes in Cross Flow, Advances in Heat
Transfer. Academic, New York, 1972.
References [25] J.G. Georgiadis, I. Catton, Dispersion in cellular thermal convection in porous
layers, Int. J. Heat Mass Transfer 31 (1988) 1081e1091.
[1] V.R. Voller, C. Prakash, A fixed grid numerical modelling methodology for [26] V.V. Calmidi, Transport phenomena in high porosity fibrous metal foams, Ph.
convectionediffusion mushy region phase-change problems, Int. J. Heat Mass D thesis, University of Colorado, 1998.
Transfer 30 (1987) 1709e1719. [27] D.L. Sun, Z.G. Qu, Y.L. He, W.Q. Tao, An efficient segregated algorithm for
[2] V.R. Voller, J.B. Swenson, C. Paola, An analytical solution for a Stefan problem incompressible fluid flow and heat transfer problems e IDEAL (inner doubly
with variable latent heat, Int. J. Heat Mass Transfer 47 (2004) 5387e5390. iterative efficient algorithm for linked equations) part I: mathematical
[3] T.J. Scanlon, M.T. Stickland, A numerical analysis of buoyancy-driven melting formulation and solution procedure, Numer. Heat Transfer, Part B. 53 (2008)
and freezing, Int. J. Heat Mass Transfer 47 (2004) 429e436. 1e17.

You might also like