Está en la página 1de 8

The effect of pore size and porosity on thermal management

performance of phase change material inltrated microcellular metal


foams
Sriharsha S. Sundarram, Wei Li
*
Department of Mechanical Engineering, The University of Texas at Austin, Austin, TX 78712, USA
h i g h l i g h t s
v Pore size and porosity of phase change material-microcellular metal foam were investigated.
v A smaller pore size results in a lower temperature at the heat source for a longer period of time.
v The effects were more pronounced at high heating and low cooling conditions.
v Net thermal conductivity doubled by reducing the pore size from 100 mm to 25 mm.
a r t i c l e i n f o
Article history:
Received 6 September 2013
Accepted 30 November 2013
Available online 10 December 2013
Keywords:
Phase change material
Microcellular metal foam
Thermal management system
Microelectronics cooling
Solar energy storage
Waste heat recovery
a b s t r a c t
The effect of pore size and porosity on the performance of phase change material (PCM) inltrated metal
foams, especially when the pore size reduces to less than 100 mm, is investigated in this study. A three
dimensional nite element model was developed to consider both the metal and PCM domains, with
heat exchange between them. The pore size and porosity effects were studied along with other system
variables including heat generation and dissipation of the PCM-based thermal management system. It is
shown that both porosity and pore size have strong effects on the heating of PCM. At a xed porosity, a
smaller pore size results in a lower temperature at the heat source for a longer period of time. The effects
of pore size and porosity were more pronounced at high heat generation and low convective cooling
conditions, representing the situation of portable electronics. There is an optimal porosity for the PCM-
metal foam system; however, the optimal value only occurs at high cooling conditions. The net effective
thermal conductivity of a PCM-microcellular metal foam system could be doubled by reducing the pore
size from 100 mm to 25 mm.
2013 Elsevier Ltd. All rights reserved.
1. Introduction
Thermal management systems based on latent heat storage of
phase change materials (PCMs) can be widely used in a variety of
applications, including microelectronics cooling, solar energy
storage, and waste heat recovery [1e5]. The PCM absorb a sub-
stantial amount of heat during phase transformation while keeping
the temperature nearly constant. This property is highly useful, e.g.,
to suppress temperature spikes due to the intermittent high power
demand in portable electronics and to store a large amount of solar
energy without signicantly increasing the system temperature.
However, PCMs in general suffer from an inherent low thermal
conductivity, which can result in slow heat dissipation and an
uneven melt front [5e8]. To energy storage systems, this means a
slow charging and discharging rate, thus a low system efciency.
For microelectronics cooling, it can cause hot spots and signicantly
shorten the chip life.
A number of techniques have been developed to enhance the
thermal conductivity of PCMs, including adding highly conductive
nanollers [6,9e12], embedding internal ns [13,14], and inl-
trating the PCM in graphitic [8,15,16] and metal foams [5,7,17e19].
Inltrating PCMs in metal foams recently attracted more attention
due to the high isotropic thermal conductivity provided by the
metal struts, which form a continuous network that spreads the
heat more rapidly throughout the PCM matrix.
The performance of metal foam-based PCM systems has been
studied both experimentally and numerically. Most of the existing
research focuses on demonstrating the performance improvement
over pure PCM-based thermal management systems and the free-
and forced-convection heat transfer phenomena inside the porous
* Corresponding author. Tel.: 1 512 471 7174.
E-mail address: weiwli@austin.utexas.edu (W. Li).
Contents lists available at ScienceDirect
Applied Thermal Engineering
j ournal homepage: www. el sevi er. com/ l ocat e/ apt hermeng
1359-4311/$ e see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.applthermaleng.2013.11.072
Applied Thermal Engineering 64 (2014) 147e154
media [5,20]. Few have considered the effect of foam morphology
including pore size and porosity. In addition, the metal foams
studied were all limited by commercial availability, with pore sizes
ranging from 0.5 to 5 mm (5e40 PPI). Hong and Herling [21,22]
experimentally studied the effect of surface area density on the
performance of parafn inltrated aluminumfoams with pore sizes
from 500 mm to 2 mm. The results suggested that heating and
cooling time of the system was affected by pore size at a xed
porosity. A smaller pore size resulted in a higher effective thermal
conductivity of the overall system. Lafdi et al. [17] also conducted
an experimental study with parafn-inltrated aluminum foams
and found that both pore size and porosity affected the perfor-
mance of the system. The aluminum foams used had a pore size
between 600 mm and 5 mm. Tian and Zhao [19] performed similar
experiments with copper foams with pore sizes from 800 mm to
2.5 mm. It was found, however, that the pore size effect was
insignicant on the melting time of inltrated PCM. Numerical
models were also developed to predict the temperature prole of
PCM metal foam systems [18,19]. These models had an origin in
Boomsma and Poulikakos [23], where the effective thermal con-
ductivity (k
e
) of an inltrated porous metal foam was estimated
based on a tetrakaidecahedron pore model. The effective k
e
was
obtained by dividing a unit cell of the pore model into layers and
applying the well-known parallel and series average models within
and between the layers according to volume fractions of foam lig-
aments and inltrated media. Although it contains the geometric
information of a pore, the Boomsma and Poulikakos model does not
consider the heat exchange between the foam ligaments and
inltrated PCM media.
With the advancement of fabrication techniques for micro-
cellular metal foams [24], the effect of pore size and porosity be-
comes more interesting due to the extremely large surface area
enabled by metal foams with pore sizes under 100 mm. It is also
interesting to understand the performance of these microcellular
metal foam PCM systems under various heating and cooling con-
ditions in order to design an optimal thermal management system.
The purpose of this study is to investigate through a modeling
approach the pore size and porosity effects of PCM-metal foam
systems, especially when the pore size is reduced to a few micro-
meters. A three dimensional (3D) nite element model is devel-
oped to represent both the metal foam and PCM domains, and
therein the heat exchange between them. The pore size and
porosity effects are studied along with system variables including
heat generation and dissipation to cover various operating condi-
tions of the PCM thermal management system. The effect of con-
vection due to molten PCM ow inside the porous structure is also
evaluated.
2. Modeling method
2.1. Heat transfer mechanisms
The heat transfer mechanisms considered in this study include
heat conduction though the metal ligaments and PCM, phase
change of PCM, heat convection in molten PCM, and convective
cooling on the external top surface of the PCM-metal foam system.
The conduction heat transfer is represented in Eq. (1),
rC
p
vT
vt
= V$(KVT)
_
q (1)
where r is the density of the material, C
p
is the specic heat ca-
pacity, T is the temperature, t is time, K is the thermal conductivity,
and _ q is the heat generation rate. The phase change process of PCM
can be represented using Eq. (2),
K
s
VT
s
K
l
VT
l
= rL
dX
dt
(2)
where L is the latent heat of fusion, X is the position of the melting
interface, and s and l stand for solid and liquid states of the PCM. In
order to account for the phase change behavior in Eq. (1), the
specic heat of PCM, C
p(PCM)
, can be dened over different tem-
perature ranges, as shown in Eq. (3),
C
p(PCM)
=
8
>
>
<
>
>
:
C
ps
; T _ T
m
C
p

L
DT
; T
m
< T < T
m
DT
C
pl
; T _ T
m
DT
(3)
where C
ps
and C
pl
are the specic heats of the PCM in solid and
liquid state, respectively, T
m
is the temperature when melting
starts, and DT is the temperature range over which PCM melts.
The convective heat transfer within molten PCM is represented
using Eq. (4),
rC
p
vT
vt
rC
p
u$VT = V$(KVT) (4)
where u is the velocity eld of the molten PCM. The ow of molten
PCM due to density and gravity can be represented using Eq. (5),
r
vu
vt
r(u$V)u = V$
h
pI m(Vu) m(Vu)
T
i
grb

T T
ref

(5)
where I is the identity matrix, g is the standard gravitational ac-
celeration, r and b are density and coefcient of thermal expansion
of the PCM, respectively, and T
ref
is the reference temperature. The
non-linear, transient heat transfer Eq. (1)e(5) need to be solved
using a numerical procedure to determine the temperature prole
in the PCM inltrated metal foam.
2.2. The nite element model
The metal foam in this study was considered to have a face-
centered-cubic (FCC) structure. The morphology of metal foams is
dependent on the material and fabrication technique [25]. The use
of polymer sphere template is one of the techniques to fabricate
small pore size metal foams [26], which results in metal foams with
pores arranged in the FCC arrangement. A geometric model of the
metal foam was created by deducting 1/8 of a sphere at the eight
corners and 1/2 sphere from the six faces of a solid cube. A unit cell
of the metal foam in FCC conguration is shown in Fig. 1(a). The
unit cell structure of the phase change material was obtained by
taking an inverse of the metal foam structure, as shown in Fig. 1(b).
The two unit cells can be overlaid to form a single unit cell con-
taining both metal and PCM, which can then be replicated to forma
PCM-metal foam model of various dimensions.
Finite element simulation of transport properties of metal foams
has been previously conducted with a body-centered-cubic (BCC)
packing structure [27]. At a sphere radius of 0.5 of the unit cell size,
the BCC structure ceases to be completely open-celled, and the
corresponding porosity at this condition is 0.94. BCC models with
porosity lower than 94% over predicts experimental results,
because of the larger volume fraction of solid in the closed-cell
model. The FCC packing structure used in this study was gener-
ated with a constant pore size, such that models with porosities
above 75% remained an open-celled structure. The same model
structure has been used to predict the thermal conductivity of
micro- and nano-cellular polymer foams with satisfactory results
[28].
S.S. Sundarram, W. Li / Applied Thermal Engineering 64 (2014) 147e154 148
A FCC model with two overlaid unit cells replicated in one di-
rection (z-axis) is shown in Fig. 2. Metal slabs were added to the
two ends to represent the housing in which the PCM system was
enclosed. A thickness equal to one-tenth the unit cell size was
arbitrarily chosen for the metal housing. Another metal slab with
half the thickness was attached to one end of the PCM unit to
represent the heat source. Assuming the PCM system is at with a
thickness signicantly smaller than the lateral size, the congura-
tion shown in Fig. 2 was used to model a one-dimensional heat
transfer process through the thickness of the system. Because of
symmetry, the adiabatic condition was imposed on the four sides of
the model. A convective boundary condition was applied to the top
surface of the PCM system to represent a cooling situation by
natural or force air convection.
A no-slip boundary condition was enforced at the PCMand foam
ligament interfaces. A contact pair was dened to allow heat ex-
change between the two phases. A continuity constraint was
enforced so that the heat ux and temperature across the interface
was continuous,
n
PCM
$(KVT)
PCM
= n
metal
$(KVT)
metal
T
PCM
= T
metal
(6)
where n
PCM
and n
metal
are the normal vectors of the PCMand metal
surfaces, respectively.
The materials chosen in this study for metal foam and phase
change material were aluminum and parafn wax. Their properties
are listed in Table 1. The modied specic heat dened in Eq. (3)
was implemented using a Heaviside step function to account for
the latent heat of parafn. A temperature dependent viscosity was
used to account for the phase change from solid to liquid. Both of
these two properties are plotted in Fig. 3 as functions of tempera-
ture. A nite element model of the PCM-metal foam system was
developed using the commercial nite element analysis package
COMSOL. An extremely ne mesh with a minimum element size of
0.57 mm was generated for all the simulation models. This is the
Fig. 1. The solid model of a unit cell of (a) metal foamwith face centered cubic pore conguration and (b) the phase change material obtained by taking an inverse of the metal foam
structure.
Fig. 2. Solid model of phase change material inltrated metal foam system along with imposed boundary conditions: (a) a schematic representation of a PCM system used for
microelectronics cooling; (b) an element taken through the thickness direction for modeling.
Table 1
Properties of aluminum and parafn wax.
Aluminum Parafn wax
Density (Kg/m
3
) 2800 780
Specic heat (J/Kg-K) 910 2500 (solid and liquid)
Thermal conductivity (W/m-K) 237 0.20
Latent heat (KJ/Kg) e 180
Melting range (K) e 321e335
Viscosity (Pa S) e 0.024 (liquid)
S.S. Sundarram, W. Li / Applied Thermal Engineering 64 (2014) 147e154 149
nest mesh size possible in COMSOL. Models with high porosity
(>90%) could only be meshed using this mesh size because of the
presence of narrow struts. This mesh size was used for all the cases
to maintain the consistency of simulation results. A nonlinear time
dependent solver with backward differentiation and automatic
time stepping was employed for computation.
The developed model was validated with experimental data and
used to study the effects of pore size and porosity under various
heat generation and cooling conditions. The parameters used in the
study are shown in Table 2. The heat generation rates were selected
to represent typical low and high power generation situations in
portable electronics. The convective heat transfer rates were
selected to represent free and forced convection by air on metal
surfaces. The temperature at the center of the heat source, repre-
senting the average chip temperature, was used as a response
variable for analysis. The contribution of convective heat transfer
inside the molten PCM was also examined.
3. Results and discussion
3.1. Model validation
The developed model was validated with experimental data
found in literature. A copper foam with 2.5 mm pore size and 95%
porosity was used for comparison with data adapted from Zhao
et al. [5]. The thickness of the model was 25 mm. The power gen-
eration density was 3.2e5 W/m
3
. A copper foam with 635 mm pore
size and 90% porosity was used for comparison with data adapted
fromLi et al. [18]. The thickness of the model was 22.5 mm, and the
power density for heat generation was 5e7 W/m
3
. Both models
were insulated at the top surface. The comparisons are shown in
Fig. 4. As can be seen, the predicted temperature responses match
the experimental measurements closely.
Before the model prediction was compared to experimental
data, the effect of the number of unit cells in the x, y and z di-
rection was examined. It was found that varying the numbers of
unit cells in the x and y directions did not change the simulation
results. Hence, only one unit cell was used in the x and y directions
for all the cases in this study. The number of unit cells in the z
direction was determined based on the thickness of the PCM
system.
The thermal response of the PCM system with and without the
metal foam was also examined. A model with 2 mm pore size, 94%
porosity, and one unit cell on the z-axis was used for the case with
metal foam. A power generation density of 1.25e6 W/m
3
and an
adiabatic top surface condition were applied to both cases. The
temperature proles across the mid plane along z-axis for the two
cases during PCM melting are compared in Fig. 5. It is seen that the
temperature distribution in PCM is more uniform along the thick-
ness direction when metal foam is used. For pure PCM, melting is
close to the heat source, leading to a sharp increase in local
temperature.
3.2. The effect of convective heat transfer inside molten PCM
The above PCM-metal foam model was used to study the effect
of convective heat transfer in molten PCM. The transient temper-
ature response at the center of the power source with and without
the inclusion of the convection mechanism in the molten PCM is
shown in Fig. 6. It is seen that the temperature difference between
the two cases is insignicant, implying that convection could be
neglected in the model. The velocity of PCM movement during
melting across the mid plane has been found to be on the order of
10
8
m/s. This lowvelocity results in a Grashof number much lower
than the cutoff value (10
3
) for convection to be signicant. The high
viscosity of molten PCM (0.0269 Pa s) combined with its low co-
efcient of thermal expansion (3.085 10
4
K
1
) is the reason for
such a low velocity. As the pore size reduces further, the uid ve-
locity inside the porous structure will be even lower because of
increased tortuosity. Therefore, it is reasonable to assume that the
effect of convective heat transfer will be insignicantly small in
PCM inltrated microcellular metal foams; and only conduction
though the metal foam and PCM needs to be considered in the
models.
Fig. 3. Specic heat of the phase change material represented as a Heaviside step and
temperature dependent viscosity function to account for the phase change process.
Table 2
Parameters for studying pore size and porosity effect.
Parameter Value
Pore size (mm) 5, 10, 25, 50, 100
Porosity 75, 80, 84, 88, 94
Heat generation (W/m
3
) 2.5e8, 3.2e5, 5e7, 12.5e8
Convective heat transfer coefcient (W/m
2
-K) 1, 10
Fig. 4. Predicted temperature compared to experimental data for copper foams.
S.S. Sundarram, W. Li / Applied Thermal Engineering 64 (2014) 147e154 150
3.3. Effect of pore size
Models with pore size varying from 5 to 100 mm at a porosity of
94% were generated to study the pore size effect. The thickness of
the system was xed at 126 mm. A heating rate of 2.5e8 W/m
3
was
applied to the heat source and a convective cooling coefcient of
10 W/m
2
-K was applied to the top surface of the system. The
transient temperature response at the center of the power source is
shown in Fig. 7. It is seen that the temperature rises similarly for all
pore size cases until the PCM starts to melt at 321 K. A smaller pore
size results in a longer phase transition period. For example, the
transition zone lasts for 190 s for 5 mm pore size comparing to 120 s
for 50 mm pore size. After the PCM completes melting, the tem-
perature continues to rise. The temperatures for different pore sizes
eventually all converged to the same steady-state temperature.
However, before the equilibrium is reached, the temperatures are
dramatically different at any given time. For example, the temper-
ature at t =220 s is 335 K for the 5 mmpore size, while it is 378 K for
the 50 mm pore size. This temperature difference is signicant for
PCM-metal foam applications. For example, in microelectronic
cooling applications, the failure rate of IC chips is dependent on its
temperature with an Arrhenius relationship, as shown below [29],
l = Ce
Ea
kT
(7)
where l is the failure rate, E
a
is an activation energy, k is Boltz-
manns constant, T is the operating temperature, and C is a con-
stant. It can be found from the above equation that reducing the
operating temperature by 10

would reduce the failure rate by half.


Hence, the lower temperature provided by smaller pore sized metal
foams is highly benecial to extending the lifetime of IC chips.
In order to understand the temperature difference in different
metal foams, the temperature proles of PCMin metal foams with a
pore size of 50 mm and 10 mm are examined at t = 40 s, as shown in
Fig. 8(a) and (b), respectively. The size of the models are different
because only one unit cell was used in the x and y directions. It is
observed that melting is less uniform in the foam of 50 mm pore
size, with areas of un-molten PCM surrounded by molten ones. The
melt front is also less uniform and lags behind the one of the 10 mm
case. This difference in the melting behavior could be explained by
the surface area density of foams. Surface area density increases as
the pore size is reduced at a xed porosity. For the 10 mm pore size
case, the surface area density is 3.62 mm
2
/mm
3
, whereas for the
Fig. 5. Temperature prole across mid-plane along z-axis for pure PCM and PCM inltrated metal foam (temperature in K and the region with wireframe indicates metal foam).
Fig. 6. Temperature response at center of power source in PCM inltrated metal foam
showing effect of convective heat transfer in PCM (pore size: 2 mm, porosity: 94%).
Fig. 7. Temperature response at center of power source in PCM inltrated metal foam
with pore size from 5 to 100 mm (porosity: 94%).
S.S. Sundarram, W. Li / Applied Thermal Engineering 64 (2014) 147e154 151
50 mm pore size case it is 0.71 mm
2
/mm
3
. The surface area densities
were obtained fromthe geometrical models in COMSOL. The higher
surface area in the foam of a smaller pore size results in a larger
amount of PCM in contact with the metal foam, hence more heat
dissipation into PCM and less heat accumulation at the heat source.
At any given instance of time, the temperature at the heat source
will be lower when a smaller pore sized metal foam is used.
3.4. Effect of porosity
Models with porosity varying from75 to 94% at a xed pore size
of 25 mm were generated to study the effect of porosity. The
maximum porosity that could be modeled in this study was 94%
due to the software limitation. At porosities below 75% the cells
become isolated and cannot be used for PCM inltration. Four unit
cells along the z-axis were used in the model. A heating rate of
2.5e8 W/m
3
and convective cooling coefcient of 10 W/m
2
-K were
applied. The transient temperature response at the center of the
power source is shown in Fig. 9 for the rst 10 s of heating. In
general a lower porosity resulted in a lower steady-state temper-
ature. At a lower porosity, the volume fraction of metal is higher.
The increased fraction of metal helps conduct heat away rapidly,
leading to a lower steady-state temperature. However, it should be
noted that a lower foam porosity would result in increased weight
and a lower heat storage capacity of the overall system due to the
reduced amount of phase change material in the system.
3.5. Effect of pore size and porosity with varying heating and
cooling conditions
PCM-metal foam systems may be used under various heating
and cooling conditions. Therefore, the effects of pore size and
porosity need to be examined in conjunction with operation con-
ditions of the system. Fig. 10 shows the predicted temperatures
under different pore size, porosity, heating, and cooling conditions.
Two different pore sizes, 25 mm and 100 mm, were studied with
porosities ranging from75% to 94%. Two heat generation conditions
were set at 2.5e8 W/m
3
and 12.5e8 W/m
3
. Two cooling conditions
were set at 1 W/m
2
-K and 10 W/m
2
-K. The temperature at the
center of the heat source at t = 10 s was used for comparison.
It is seen from Fig. 10(a) that the pore size and porosity effects
are almost indistinguishable under low cooling and low heat gen-
eration conditions. At the high convective cooling rate, the effect of
porosity can be clearly seen, as shown in Fig. 10(b) and (d), while
the pore size effects are not pronounced. Under high cooling but
low heating conditions, the temperature remains almost constant
until the porosity reaches 88%. At high cooling and high heat gen-
eration conditions, the temperature rst reduces as the porosity
increase, and increases after the porosity is above 84%. Both the
pore size and porosity effects are signicant under lowcooling and
high heating conditions, as seen in Fig. 10(c). On average, whenpore
size is reduced from 100 mm to 25 mm, the temperature is reduced
by 10 K; when porosity is increased from 75% to 94%, the temper-
ature is reduced by 15 K. In portable electronic devices, natural
convection on the external surface is the major mechanism of heat
dissipation, while power demand could be high [30]. This situation
is represented in Fig. 10(c), where the benet of small pore size and
high porosity is clearly seen.
It has been suggested previously that PCM inltrated metal
foam with a porosity of 85% would yield the best thermal perfor-
mance [5,19]. However, the simulation results in this study show
that it could be true only when the convective cooling rate is high.
Heat conduction through the PCM-metal foam system has two
components. At a lower porosity, conduction through the metal
ligaments is dominant because of its high thermal conductivity. At
85% porosity, heat dissipation by convection and heat absorption by
the phase change material are balanced, resulting in the lowest
temperature. This behavior is not observed at high heat generation
and low convective cooling conditions.
3.6. The effective thermal conductivity of PCM-metal foam system
The effective thermal conductivity of the PCM inltrated metal
foams was determined for pore sizes of 25 mm and 100 mm at
porosity of 88% across the thickness of PCM-metal foam system as
K
eff
=
q
(vT=vz)
(8)
where K
eff
is the effective thermal conductivity and q is the heat
ux. The temperature values at t = 10 s for the four cases shown in
Fig. 10 were used to evaluate Eq. (8). The effective thermal con-
ductivity values are plotted as a function of temperature in Fig. 11.
Fig. 8. Melt front in PCM inltrated metal foamwith pore size (a) 50 mm and (b) 10 mm
(region with wireframe indicates metal foam).
Fig. 9. Temperature response at center of power source in PCM inltrated metal foam
with varying porosity and pore size 25 mm.
S.S. Sundarram, W. Li / Applied Thermal Engineering 64 (2014) 147e154 152
The horizontal axis represents the temperature at the center of the
heat source.
It is seen that at a xed porosity, the thermal conductivity is
dependent on pore size. A smaller pore size results in a higher
thermal conductivity. On average the thermal conductivity more
than doubled, from 5 W/m-K to 12 W/m-K when the pore size
reduced from100 mm to 25 mm. Therefore, a smaller pore size PCM-
metal foam system is benecial as it results in a higher effective
thermal conductivity for faster heat dissipation, which leads to a
lower temperature at the heat source.
4. Conclusions
The effect of pore size and porosity of phase change material
inltrated microcellular metal foams was studied in this research. A
nite element based model was developed to represent the metal
foam-PCM system and the model was validated with experimental
data. It is shown that both pore size and porosity strongly affect the
performance of the PCM-metal foam thermal management system.
At a xed porosity, a smaller pore size leads to a longer time period
that the temperature of the heat source remains nearly constant.
The increase in this time period is attributed to the higher surface
area of smaller pore size metal foams and its high thermal con-
ductivity, which helps dissipate the heat rapidly away fromthe heat
source. The effects of pore size and porosity also depend on the heat
generation and cooling conditions of the system. The pore size and
porosity effect is more pronounced under high heat generation and
low convective cooling conditions which represent the situation of
portable electronics. At these conditions, the net effective thermal
conductivity could be doubled by reducing the pore size from
100 mm to 25 mm. There seems an optimal porosity for the PCM-
metal foam system; however, the optimal value only occurs at
high cooling conditions. In order to design an optimal thermal
management system, the pore size and porosity, as well as the
heating and cooling conditions need to be considered
simultaneously.
References
[1] A. Hoshi, D.R. Mills, A. Bittar, T.S. Saitoh, Screening of high melting point phase
change materials (PCM) in solar thermal concentrating technology based on
CLFR, Sol. Energy 79 (3) (2005) 332e339.
[2] A. Sharma, V.V. Tyagi, C.R. Chen, D. Buddhi, Review on thermal energy storage
with phase change materials and applications, Renew. Sustain. Energy Rev. 13
(2) (2009) 318e345.
Fig. 10. Temperature response as a function of metal foam pore size and porosity for different heat generation and convective cooling rates (t = 10 s).
Fig. 11. Effective thermal conductivity of PCM inltrated metal foam as a function of
temperature for varying pore sizes (porosity: 88%).
S.S. Sundarram, W. Li / Applied Thermal Engineering 64 (2014) 147e154 153
[3] V.V. Tyagi, D. Buddhi, PCM thermal storage in buildings: a state of art, Renew.
Sustain. Energy Rev. 11 (6) (2007) 1146e1166.
[4] B. Zalba, J.M. Marn, L.F. Cabeza, H. Mehling, Review on thermal energy storage
with phase change: materials, heat transfer analysis and applications, Appl.
Therm. Eng. 23 (3) (2003) 251e283.
[5] C.Y. Zhao, W. Lu, Y. Tian, Heat transfer enhancement for thermal energy
storage using metal foams embedded within phase change materials (PCMs),
Sol. Energy 84 (8) (2010) 1402e1412.
[6] A.S. Fleischer, K. Chintakrinda, R. Weinstein, C.A. Bessel, Transient thermal
management using phase change materials with embedded graphite nano-
bers for systems with high power requirements, in: 11th Intersociety Con-
ference on Thermal and Thermomechanical Phenomena in Electronic Systems,
2008.
[7] P.H. Vadwala, Thermal Energy Storage in Copper Foams Filled with Parafn
Wax, University of Toronto, Toronto, 2011.
[8] Y. Zhong, Q. Guo, S. Li, J. Shi, L. Liu, Heat transfer enhancement of parafn wax
using graphite foam for thermal energy storage, Sol. Energy Mater. Sol. Cells
94 (6) (2010) 1011e1014.
[9] A. Elgafy, K. Lafdi, Effect of carbon nanober additives on thermal behavior of
phase change materials, Carbon 43 (15) (2005) 3067e3074.
[10] J. Fukai, M. Kanou, Y. Kodama, O. Miyatake, Thermal conductivity enhance-
ment of energy storage media using carbon bers, Energy Convers. Manag. 41
(14) (2000) 1543e1556.
[11] S. Kim, L.T. Drzal, High latent heat storage and high thermal conductive phase
change materials using exfoliated graphite nanoplatelets, Sol. Energy Mater.
Sol. Cells 93 (1) (2009) 136e142.
[12] W. Wang, X. Yang, Y. Fang, J. Ding, J. Yan, Enhanced thermal conductivity and
thermal performance of form-stable composite phase change materials by
using b-aluminum nitride, Appl. Energy 86 (7e8) (2009) 1196e1200.
[13] P. Lamberg, K. Sirn, Analytical model for melting in a semi-innite PCM
storage with an internal n, Heat Mass Transfer 39 (2002) 167e176.
[14] V. Shatikian, G. Ziskind, R. Letan, Numerical investigation of a PCM-based heat
sink with internal ns: constant heat ux, Int. J. Heat Mass Transfer 51 (5e6)
(2008) 1488e1493.
[15] K. Lafdi, O. Mesalhy, A. Elgafy, Graphite foams inltrated with phase change
materials as alternative materials for space and terrestrial thermal energy
storage applications, Carbon 46 (1) (2008) 159e168.
[16] O. Mesalhy, K. Lafdi, A. Elgafy, Carbon foam matrices saturated with PCM for
thermal protection purposes, Carbon 44 (10) (2006) 2080e2088.
[17] K. Lafdi, O. Mesalhy, S. Shaikh, Experimental study on the inuence of foam
porosity and pore size on the melting of phase change materials, J. Appl. Phys.
102 (8) (2007) 083549e083549-6.
[18] W.Q. Li, Z.G. Qu, Y.L. He, W.Q. Tao, Experimental and numerical studies on
melting phase change heat transfer in open-cell metallic foams lled with
parafn, Appl. Therm. Eng. 37 (2012) 1e9.
[19] Y. Tian, C.Y. Zhao, A numerical investigation of heat transfer in phase change
materials (PCMs) embedded in porous metals, Energy 36 (9) (2011) 5539e
5546.
[20] A. Siahpush, J. OBrien, J. Crepeau, Phase change heat transfer enhancement
using copper porous foam, J. Heat Transfer 130 (8) (2008) 082301-1e
082301-11, http://dx.doi.org/10.1115/1.2928010.
[21] S.T. Hong, D.R. Herling, Effects of surface area density of aluminum foams on
thermal conductivity of aluminum foam-phase change material composites,
Adv. Eng. Mater. 9 (7) (2007) 554e557.
[22] S.-T. Hong, D.R. Herling, Open-cell aluminum foams lled with phase change
materials as compact heat sinks, Scr. Mater. 55 (10) (2006) 887e890.
[23] K. Boomsma, D. Poulikakos, On the effective thermal conductivity of a three-
dimensionally structured uid-saturated metal foam, Int. J. Heat Mass
Transfer 44 (4) (2001) 827e836.
[24] S.S. Sundarram, W. Jiang, W. Li, Fabrication of small pore-size nickel foams
using electroless plating of solid-state foamed immiscible polymer blends,
J. Manuf. Sci. Eng. (2013), http://dx.doi.org/10.1115/1.4025418.
[25] J. Banhart, Manufacture, characterisation and application of cellular metals
and metal foams, Prog. Mater. Sci. 46 (6) (2001) 559e632.
[26] H. Zhang, X. Yu, P.V. Braun, Three-dimensional bicontinuous ultrafast-
charge and -discharge bulk battery electrodes, Nat. Nanotechnol. 6 (5)
(2011) 277e281.
[27] S. Krishnan, J.Y. Murthy, S.V. Garimella, Direct simulation of transport in open-
cell metal foam, J. Heat Transfer 128 (8) (2006) 793e799.
[28] S.S. Sundarram, W. Li, On thermal conductivity of micro- and nanocellular
polymer foams, Polym. Eng. Sci. 53 (9) (2013) 1901e1909.
[29] A. Vassighi, M. Sachdev, Thermal and Power Management of Integrated Cir-
cuits, 2006 ed., Springer, New York, NY, 2006, p. 192.
[30] E.M. Alawadhi, C.H. Amon, PCM thermal control unit for portable electronic
devices: experimental and numerical studies, IEEE Trans. Compon. Packag.
Technol. 26 (1) (2003) 116e125.
S.S. Sundarram, W. Li / Applied Thermal Engineering 64 (2014) 147e154 154

También podría gustarte