Вы находитесь на странице: 1из 8

JOURNAL OF THERMOPHYSICS AND HEAT TRANSFER

Vol. 29, No. 4, OctoberDecember 2015

Modeling Thermal Storage in Wax-Impregnated


Foams with a Pore-Scale Submodel
Galen R. Jackson, Kyle C. Smith, Patrick C. McCarthy, and Timothy S. Fisher
Purdue University, West Lafayette, Indiana 47907

Downloaded by PURDUE UNIVERSITY on May 26, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.T4523

DOI: 10.2514/1.T4523
Calculations of heat transfer through porous foams filled with phase-change materials are often performed using
either an effective medium approach or a coupled two-temperature model. These models are generally applied to
simulate heating processes that occur over a few hours, thus allowing for an assumption of effective melting front in
which both liquid and solid phases are present or clear delineation of liquid and solid phases. However, in the instance
of a rapidly changing heating process, a melting front cannot be assumed due to the possibility that layers of phases
exist. A model has been developed to allow for variable phase sections within the foam without requiring an
assumption of local thermal equilibrium, in which the temperature of the foam and phase-change materials are equal
to each other at a set point. By assuming spherical pores with microscopic temperature variations therein, the model
allows for the existence of multiple phases within a single foam pore while also integrating an effective medium model
of the overall foam/wax system. The overall model is compared and validated with experimental results and models
from literature. Following validation, cases are assessed under various time scales to demonstrate the models
versatility.

Nomenclature
A
asf
cp
D
f
G
g
H
h
K
L
N
NuD;i
q
Re
r
SA
T
T i;j
t
U
V

H
r
T
x

=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=

area, m2
interfactial surface area per unit volume, m1
specific heat capacity, k J kg1 K1
diameter, m
melt fraction
interfacial thermal conductance, W m2 K1
gravitational acceleration, m s2
enthalpy, Jkg
heat transfer coefficient, W m2 K1
permeability, m2
system height, m
number of
interstitial Nusselt number
heat flux, W m2
Reynolds number
radius, m
surface area, m2
temperature, K
temperature of node i, j, K
time, s
velocity, ms
volume, m3
thermal diffusivity, m2 s1
thermal expansion coefficient, K1
latent heat, Jkg
distance between nodes in the radial direction, m
temperature difference, K
distance between nodes in the x direction, m

=
=
=
=
=
=
=

distance between nodes in the y direction, m


porosity
thermal conductivity, W m1 K1
heating/cooling period, s
kinematic viscosity, m2 s1
density, kg m3
dimensionless time

Subscripts
C
eff
f
H
i
j
l
m1
m2
m, avg
P
p
s

=
=
=
=
=
=
=
=
=
=
=
=
=

cold
effective
foam
hot
horizontal foam node
vertical foam node
liquid
melt start
melt finish
melt average
phase-change materials inside pore
pore node
solid

Superscript
n

time iteration number

I.

Introduction

HERMAL management in electronic devices is vital for


both efficient operation and longevity. Because of this need,
numerous investigations in the past 30 years have used latent heat
storage for thermal management. Phase-change materials (PCMs) are
often used in thermal management systems that take advantage
of latent heat storage because PCMs maintain a near-constant
temperature during heat absorption. This form of latent heat energy
storage is especially useful in applications in which a temperature
limit is necessary, such as cooling electronic devices. An additional
advantage of latent heat energy storage is that PCMs remove the need
for moving parts, such as a cooling fan in a computer heat sink, from
inside the system. For example, latent heat energy storage can be used
to cool photovoltaics in low Earth orbit applications [1].
Ideal PCMs have high thermal conductivity, high latent heat, and
small density change during phase change. However, most common
PCMs have a low thermal conductivity. To enhance effective thermal

Presented as Paper AIAA-2014-3122 at the 11th Annual AIAA/ASME


Joint Thermophysics and Heat Transfer Conference, Atlanta, GA, 16
20 June 2014; received 6 July 2014; revision received 23 October 2014;
accepted for publication 24 October 2014; published online 30 January 2015.
Copyright 2014 by the American Institute of Aeronautics and Astronautics,
Inc. All rights reserved. Copies of this paper may be made for personal or
internal use, on condition that the copier pay the $10.00 per-copy fee to the
Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923;
include the code 1533-6808/15 and $10.00 in correspondence with the CCC.
*Research Assistant, Mechanical Engineering, 1205 West State Street.
Student Member AIAA.

Postdoctoral Research Associate, Mechanical Engineering, 1205 West


State Street.

Research Assistant, Mechanical Engineering, 1205 West State Street.

James G. Dwyer Professor, Mechanical Engineering, 1205 West State


Street. Senior Member AIAA.
812

813

JACKSON ET AL.

transport, fibers [2], fins [3], and porous structures [4] composed of
highly conductive materials have been incorporated with PCMs. Of
all the enhancement techniques, the greatest emphasis has been
placed on thermally conductive porous structures, specifically metal
foams [58] and high conductivity carbon foams [9,10]. Prior
modeling of heat transfer within these complex structures has
followed one of two methods: a one-temperature model and a twotemperature model.
The one-temperature effective medium model can be expressed
as [11]
cp;eff T

 cp;eff UT   eff T H


t
t

(1)

Downloaded by PURDUE UNIVERSITY on May 26, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.T4523

where cp;eff and eff are the effective thermal capacitance and
thermal conductivity, and H is the latent heat multiplied by the
porosity . The two-temperature model has also been reported [12]:
H P
 qP  hasf T f T P 
t
T f
 qf hasf T f T P 
1 f cf
t
P

(2)

where asf is the interfacial surface area per unit volume, h is the heat
transfer coefficient, and qf and qP are the heat flux inside the foam
and PCM, respectively. In the one-temperature approach, the system
is assumed to exist under local thermal equilibrium (LTE) using
effective properties to represent the foam and the PCM in the
diffusion equation. For a system to be considered under LTE, there
must be no temperature difference between the foam and the PCM
inside the pores (i.e., T f T P ).
The local thermal equilibrium approximation is invalidated in
situations where there are large differences between the thermal
properties of the foam and PCM [13] (such as thermal diffusivity),
where convection is significant [14], and when the system is under
rapid transient boundary conditions [15]. The validity of the LTE
approximation can be checked using the Sparrow number developed
by [15]. A large Sparrow number indicates that the LTE approximation is accurate (Sp > 100 for most applications). Increasing the
rate at which the boundary conditions change increases the Sparrow
number needed for the system to remain in LTE.
The two-temperature approach is the most common model used in
literature and separately predicts the temperatures for the foam and
the PCM at a given position. This approach can illustrate the effects of
thermal nonequilibrium and simultaneous liquid-phase convection.
However, modeling convection also requires conservation of mass
and linear momentum equations. For PCM filled foams, natural
convection is constrained and is thus deemed negligible by some twotemperature models, converting the problem to pure conduction [16].
To neglect natural convection, the system must adhere to the criteria
Raeff 104 , provided by [17], where Raeff is the effective Rayleigh
number of the porous medium [18]:
Raeff 

gKLT

one region can be approximated using the melt fraction parameter f.


The melt fraction can be used to determine whether the PCM is solid
(f  0), liquid (f  1), or melting (0 < f < 1). The melting fraction is
determined using the definition from [21]:

f

8
<

T P T m1
: T m2 T m1

T P < T m1
T m1 T P T m2
T m2 < T P

(4)

where T P is the temperature of the PCM inside a pore, T m1 is the


starting melting temperature, and T m2 is the ending melting
temperature.
Most applications of previous research related to PCMs and latent
energy storage have time scales of hours, and the bulk of research for
this topic either uses or assumes quasi-steady state, also known as
constant boundary conditions. This is the case for all prior studies
mentioned in this section. The goal of this present work is to create a
model capable of calculating thermal storage under rapidly transient
conditions. This requirement prevents the use of the two-temperature
model because the assumption that all pores are isothermal becomes
invalid. Consequently, this work proposes a new model that
discretizes a pore into nodes, with each node capable of determining
whether it holds a liquid or solid phase.

II.

Model

To represent the transient nature of foam saturated with PCM, a twoscale model was developed that calculated the temperature profile of
the pores as a submodel inside an effective medium one-temperature
foam model. With short enough heating/cooling time scales, it is
possible that PCM inside a pore will not be completely melted before
solidification begins again. This phenomenon produces multiple phase
layers inside one pore. Figure 1 shows a diagram comparing a pore
under a slow heating/cooling cycle to a fast heating/cooling cycle.
As can be observed in Fig. 1, the conditions surrounding the fast
heating/cooling cycle of the pore do not heat it long enough to fully
melt the PCM before the cooling cycle begins again, producing
multiple layers of liquid and solid-phase PCM material.
The proposed model separates the foam/PCM domain into two
separate temperature profile calculations. Because of the nonisothermal phase change in manufactured PCMs, a diffuse interface
approach is employed here for the melting and solidification processes.
This approach is similar to the enthalpy method of [22] that implements
a linear approximation of the enthalpy profile between melting
temperatures of the PCM. Figure 2 depicts the enthalpy method based
on a PCM with a single temperature melting range.
Using the assumption that the enthalpy changes linearly with
temperature, an effective specific heat is approximated by taking the
derivative of the enthalpytemperature profile over the melting
temperature range. For a PCM with a latent heat H and temperature
melt range between T m1 and T m2 , an effective specific heat capacity is

(3)

where K is the permeability, L is the height of the system, is the


thermal diffusivity of the PCM, is the thermal expansion coefficient
of the PCM, and is the kinematic viscosity of the PCM.
The enthalpy method is a common approach when modeling phasechange heat transfer [19]. This method determines the overall enthalpy
trend of the PCM by using the latent heat needed to change from solid
to liquid phase (or vice versa) to produce a temperatureenthalpy trend.
This can then be used to follow when the phase-change process is
isothermal. However, because of impurities in manufactured PCMs,
the phase-change process is nonisothermal and instead occurs over a
temperature range, creating a region of temperatures in which both
liquid and solid phases exist [20]. The ratio of liquid to solid phases in

Fig. 1 Diagram of a spherical pore undergoing either slow or fast


heating/cooling cycles. T0 is the pore surface temperature, and Tm is the
melting temperature.

814

JACKSON ET AL.

eff Af

n1
n1
n1
n1
T i1;j
T n1
T i1;j
T n1
T i;j1
T n1
T n1
i;j
i;j
i;j
i;j1 T i;j



xi1;j
xi1;j
yi;j1
yi;j1

n1
N P GP SAP;end T n1
P T i;j 

n
cp f V f T n1
i;j T i;j 

tn1 tn 

(7)

for the foam node and


n1
P SAp T p1
T n1
p;end 
n1
 GP SAP;end T n1
p;end T i;j 
r
n1
T p;end
T np;end 
 P V p cp;eff
n1
t
tn 

Downloaded by PURDUE UNIVERSITY on May 26, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.T4523

Fig. 2 Enthalpytemperature relationship for an example PCM,


RT58 [4].

approximated as

cp;eff 

8
>
<

cp;s
cp;s T m;avg T m1 Hcp;l T m2 T m;avg 
T m2 T m1
>
:

cp;l

T < T m1
T m1 T T m2
T m2 < T

where T m;avg is the average of T m1 and T m2 . The enthalpy method


provides flexibility for the phase-change calculation to be
incorporated into any Cartesian, polar, or spherical coordinate
system using energy conservation:
cp;eff

T
 T
t

for the outermost pore node (note that T p;end  T p2 in Fig. 3). In the
previous equations, Af is the cross-sectional area of the foam; SAp
represents the surface area of the pore; eff is the effective thermal
conductivity of the foam; GP is the interfacial conductance between
the foam and the PCM; P is the density of the PCM; N p is the
number of pores inside node i, j of the foam; r is the change in pore
radius; T p is the temperature of the pore; T i;j is the temperature of
node i, j in the foam; and V p and V f are the volumes of the pore and
foam, respectively. The interfacial conductance term is determined
by the equation
GP 

(5)

(6)

Using Eq. (6) as a basis, the model is implemented in two separate


coordinate systems: Cartesian coordinates for the foam and spherical
coordinates for the pore. Each coordinate system was discretized into
a set number of nodes as shown in Fig. 3.
The nodal temperature distributions were solved using the finitevolume method (FVM) on the discrete grids shown in Fig. 3. Using
FVM to solve Eq. (5) at node i, j creates the equations

(8)

2NuD;i P V p;end
D2p;end SAp;end

(9)

where NuD;i is the interstitial Nusselt number. Comparing Eqs. (7)


and (8) to Eq. (2) reveals a similarity between the current model and
the two-temperature model. The main difference between the two
models is the treatment of the pore. The current model is an extension
of the two-temperature model, using interfacial conductance instead
of a heat transfer coefficient and the overall surface area of the pores
instead of the interfacial surface area.
In the current model, the thermal conductivity of the foam is the
only thermal parameter taken as an effective term when compared to
the effective medium approach [Eq. (1)] to create a term that is
homogenous. This term is taken as an effective property because the
heat conducts through both the foam and the pore inside one foam
node. The rest of the terms in Eq. (6) are the thermal properties of the
foam. For metal foams, the effective thermal conductivity method
provided by [23] is used, whereas the permeability is determined by
the method provided by [24]. The calculations can be seen in Table 1.
For pore calculations to be efficient, pores are assumed to
be spherical and located inside foam nodes. This requires the

Fig. 3 Diagram of the nodal distribution for the foam and pore submodel discretizations with the foam exposed to four different boundary conditions
(BC1, BC2, BC3, BC4).

815

JACKSON ET AL.

Table 1
Parameter
Effective thermal conductivity eff

Downloaded by PURDUE UNIVERSITY on May 26, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.T4523

Permeability K

List of important parameter correlations used in the model


Equation
p
2
eff 
2RA  RB  RC  RD 
4d
RA 
2e2  d1 e f  4 2e2 d1 ep
e 2d2
RB 
2
e 2de2 f  2e 4d
p
 e 22de  p
 2 2e
p
p
p
RC 
2d2 1 2 2e f  2 2 2e d2 1 2 2e p
2e
RD  2
e f  4 e2 p
v

  p
up
u 22 5 e3 2 2
t
8
p
d
, e  0.339.
3 4 2e e
2 D2k
K
36 1
4 1
 2  2 cos  cos1 2 1
3
3
2Dp;end
Dk 
3

assumption that the pore only interacts with one specific foam node at
a time, creating a Robin, or mixed, boundary condition between the
pore and the foam where the temperature of the surrounding foam
node is one value. In reality, this assumption is not exact because the
foam will support a nonzero temperature gradient, and some pores
could exist in two separate foam nodes at once with separate
temperatures in each node. This assumption can be alleviated by
making the length of one node of the foam larger than the diameter
of the pore. It is assumed that all the pores are represented by a
single pore size, which is an approximation of the random and
heterogeneous pore distribution of an actual pore system. For
simplicity, it is also assumed that pores have achieved the maximum
packing factor, meaning that the maximum possible number of pores,
based on the volume fraction of the foam, is packed inside the pore
node. Finally, it is assumed that the system has constant properties for
the solid and liquid phases, that natural convection is negligible
(checked by Raeff 104 condition stated earlier [17]), and that the
foam is fully saturated with PCM. In summary, the following
assumptions are made.
1) The enthalpytemperature profile of the PCM during melting is
linear.
2) Pores are spherical and all the same size.
3) The maximum possible number of pores fills the foam, based on
foam porosity.
4) The temperature of the foam surrounding a pore is constant.
5) The thermal properties for the solid and liquid phases of the
PCM are constant.
6) Natural convection is negligible.
7) The PCM fully saturates the foam.

III.

Reference
[23]

[24]

experiment. However, this layer provided insufficient insulation, and


heat loss due to natural convection was not negligible. These
boundary conditions were imposed in the model, and the resulting
temperature profiles were compared to the experimental results from
[5] (see Fig. 4). Temperature profiles were plotted at distances of 0
and yL  0.32 from the heater boundary to match the positions at
which temperatures were sampled experimentally [5]. The simulated
temperature distributions agree closely with previous experimental
measurements. Observation of time scales between experimental
results and the model indicates that the model begins and ends
melting at the proper times. The main difference between experimental and model results is the slope of the temperaturetime trend.
Close observation of the model results suggests a limitation of the
linear enthalpy assumption due to the slope of the line changing based
on the effective specific heat calculated during the phase-change
process of the PCM. Comparisons to experimental results show that
the model underestimates the difference between temperatures at
each location (i.e., the distance between the lines is smaller than the
experimental results). Potential causes for this behavior are lack of
accounting for heat losses and/or inaccuracies in the effective thermal
conductivity calculation. In this specific case, Raeff is in the order of
101 , making the conduction assumption valid.

Validation

Equations (7) and (8) were solved using an implicit integration


technique in Simulink [25]. The number of foam and pore nodes were
optimized by running a grid independence study, minimizing the
temperature change with increasing node count (i.e., the point at
which changing the number of nodes no longer affected the result)
without violating the constraint that the foam nodes must be larger
than the pore size. The tolerance of the grid independence was
selected so that the temperature change was less than 103 K. The
current model was compared to results by [5] by simulating their
experimental setup. In the test, rectangular PCM-filled copper
foam (dimensions 200 120 25 mm) was heated on the largest
surface with gravity perpendicular to the surface (heated area
200 120 mm; see BC1 of Fig. 3) with natural convection on all
other boundaries. A thin layer of acrylic was used as insulation in the

Fig. 4 Simulated temperature distributions compared to experimental


results from [5].

Downloaded by PURDUE UNIVERSITY on May 26, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.T4523

816

JACKSON ET AL.

Fig. 5

Boundary conditions for simulation setup.

Comparison to the numerical results seen from [5] shows that the
current model captures the trend of the data in the phase-change
region more accurately. It should be noted that the flat trend observed
in the numerical results of [5] is caused by a numerical assumption
that the phase change is isothermal, taken as the average temperature
in the phase-change temperature range. However, the temperature
difference issue observed in the current model is not an issue found in
the model performed in [5]. It appears that their model more
accurately estimates the thermal resistance of the system.

IV.

Results and Discussion

The simulations were run to investigate the effect pore size and
transient cycle periods had on the temperature distribution of the
foam and melt fraction inside the pores. Calculated model predictions
were expressed in nondimensional form similar to [11,26]. The
nondimensional forms for time, temperature, and pore size were
set as

 P;l
;
L2

t
t  ;

T TC
;
T 
TH TC


Dp;end
D 
y
(10)


where P;1 is the thermal diffusivity of the pore at liquid phase; H is


the system height; is the is the amount of time required to complete
one heating/cooling cycle (one period); is the dimensionless time
period; T H and T C were the hot and cold temperatures, respectively;
and y was the distance between nodes in the y direction. The model
used a constant Nusselt number, set as the upper bound of the twotemperature model Nusselt range NuD;i  5.9 from [17]. Boundary
conditions were set to a constant temperature below the PCM melting
point (T C ) at BC2 (see Fig. 3) and a pulsing temperature on the
opposite side that alternated between T H and T C . All other walls were
adiabatic. Boundary conditions and system setup were as shown in
Fig. 5. To check that the program was producing expected results, the
Table 2

Thermal properties of RT58 paraffin


wax

Property
Melt range
Latent heat
Specific heat capacity
Thermal conductivity
Density (solid)
Density (liquid)
Thermal diffusivity, approximated
Kinematic viscosity

Value
4862C
181 kJ kg1
2 kJ kg1 K1
0.2 W m1 K1
880 kg m3
770 kg m3
103 K1
32.49 mm2 s1

Fig. 6 Temperature profile along the length of the foam for D of 1.0
(left), 0.5 (middle), and 0.25 (right).

parameters used by [25] were input into the model to see if the results
were the same. After comparing the results, it was determined that the
model produced an answer with less than 5% difference from the
previous work. A full explanation of the investigation performed is
given in the Appendix to avoid confusion in the following sections.
In the model, the temperature difference was set so that
T m;avg  0.5, or in other words, the average melting point is halfway
between T H and T C . T H and T C were constrained by setting
T m2 T m1   0.1T H T C . For this investigation, the same
system size (200 120 25 mm), copper foam, and PCM (RT58)
were used as [5] to validate the model. Constant values were
set for the porosity (  0.95), foam thermal conductivity
(350 W m1 K1 ), height (L  11y), number of foam nodes
(N f  11), and number of pore nodes (N P  30). The thermal
properties for RT58 used by the model were all obtained from
RUBITHERM, with the exception of the thermal diffusivity, which
was not available (see Table 2). Instead, an approximate value for the
thermal diffusivity with scaling on the order of 106 to 103 from
waxes with similar properties was used [4,7,27]. The study used
103 K1 as a conservative estimate.
To demonstrate how a discretization of the pore might prove
useful, three different dimensionless pore sizes D were used: 1.0,
0.5, and 0.25. Models for these three pore sizes were calculated over a
range of pulse periods. Each simulation was allowed to run long
enough for the temperature profile of each foam node to continuously
repeat with the same pattern. Figure 6 shows one period of a slower

JACKSON ET AL.

817

Downloaded by PURDUE UNIVERSITY on May 26, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.T4523

Fig. 7 Melt fraction distribution inside the pore inside node 2 for D of 1.0 (left), 0.5 (middle), and 0.25 (right) at time periods a)  0.0007, and
b)  0.0021.

pulse period (  0.0021) over the three pore sizes for the foam
nodes where yH  0.1, 0.3, 0.5, and 0.7.
From Fig. 6, it was shown that pore diameter had an effect on the
overall temperature of the foam. Peak temperature was greatest for
large values of D . However, the temperature-oscillation time scale
was short enough that only the node closest to the wall where the heat
is applied (node 2) had enough time to reach the melting point.
Additionally, the transition between heating and cooling was
smoother for smaller pore diameters, which was expected because a
smaller pore diameter means less PCM volume to melt in each pore
and a greater number of pores inside one foam node.
To understand pore influence on temperature profiles, the melt
fraction distribution of the pores inside node 2 was analyzed. The
temperature at the pore wall influenced the foam temperature
significantly because it required more energy to melt the PCM than to
heat one phase. Therefore, when the melt fraction was either 0 or 1,
the foam was able to heat or cool more freely because the pore
changed temperature more quickly. The melt fraction for different
pulse widths inside the pore inside node 2 is shown in Fig. 7a for
a fast pulse period (  0.0007) and Fig. 7b for a slower pulse
period (  0.0021).
Figure 7a demonstrates an instance in which there is limited time
available for the foam to heat completely before it cools, causing the
melt fraction to be smaller than it may have otherwise been for the
slower pulse trend shown in Fig. 7b. Additionally, for D  1.0 pore
size, the pore was too big for the center of the pore to begin melting,
and PCM at the pore wall was still melting by the time the cooling
cycle began. After the cooling cycle began, the rapid cooling of the
PCM at the pore wall created a melting front seen at t  0.6 that
propagated through and got smaller as the cooling cycle continued.

Fig. 8 Contour plot showing the melt fraction change across a D  1


pore at t  0.6 with respect to dimensionless pulse period.

Because D  0.25 is a smaller pore, there was less PCM to melt, and
so the melt fraction created during the heating cycle was much larger.
Also, because there was less distance to conduct through the center of
the pore, the pore was able to melt significantly before the cooling
cycle began. Figure 7b presents the clearest image of the melting front
phenomena described by Fig. 1 at t  0.6 for D  1.0. Because
the heating cycle was longer, more of the PCM was able to melt inside
the pore. With increasing t for D  1.0, the propagation can be
observed as the melt fraction at middle of the pore at t  0.8 was
greater than at t  0.6. Even though the pore was cooling, the heat
conducted through the rest of the pore. The influence of pore size is
also shown as D increases in Fig. 7b. For t  0.6, the D  0.5
pore was only beginning to solidify after fully melting. Meanwhile,
the profile for D  0.25 is very similar to D  0.5, but the pore
was small enough that the center of the pore was able to begin
solidification as well. For both Figs. 7a and 7b, the pulse period began
with the PCM in the pore completely solid for all pore diameters. This
is a potential explanation as to why the melt fraction trends were the
same for all diameter sizes at t  0.8 because each pore had the
same amount of time to completely cool. Increasing the time
scale closer to t  1.0, pore diameters D  0.5 and D  0.25
became fully solid almost immediately, whereas D  1.0 did not
completely solidify until t  0.9. The trend seen at t  0.8 was an
unexpected phenomenon because the phase change finished so
rapidly after t  0.8 for the smaller pore sizes, whereas D  1.0
took noticeably longer to solidify. It was not anticipated that the trend
would be the same for all pore diameters, and so further investigation
would be required to determine exactly why t  0.8 is independent
of pore size.
The current model is useful for when the pulse period is small
enough that a melt fraction profile exists in the pore. However, if the
pore is completely solid or liquid for the majority of the time period,
then the districtization of the pore is unnecessary and the added
complexity of the model is no longer useful. To test the range for
which the pore scale model is appropriate, a test was run for a varying
pulse period of the D  1.0 pore size for the pore in node 2 (see
Fig. 5) at t  0.6, as shown in Fig. 8, where the contour color is the
melt fraction inside the pore.
Figure 8 gives a range of dimensionless pore period values that
provide a range of applicability for the model. Following the figure,
from  0, the pore was completely liquid until a pore period of
 0.0003 was reached, at which point the temperature of the foam
was given enough time to reach a temperature hotter than the PCMs
initial melting point. The PCM at the center of the pore did not begin
to melt until the pore pulse reached  0.0015. After this point, the
melt front shown by the fast heating/cooling cycle in Fig. 1 was seen
throughout the pore until it reached  0.0033, and the center of the
pore was fully melted before the cooling cycle began. After this point,
the pulse period transitioned to the slow heating/cooling cycle shown
in Fig. 1, where only a single melting front can be seen inside the
pore. This melting front profile continues to be observed until

818

JACKSON ET AL.

 0.0177, the cutoff point for the current model. After this
point, the pore is either all solid or all liquid, and an isothermal
temperature profile is applicable, and so there is no reason to
districtize the pores.

Downloaded by PURDUE UNIVERSITY on May 26, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.T4523

V.

Conclusions

A model was presented with the capability of calculating transient


heating of foams saturated with PCM. Heating was calculated by
discretizing both the foam and spherical pores into nodes. The model
was validated against experimental results from literature as well as
computational results for pulsed heating. The temperature profile
of the foam as well as the melt fraction profile of the pores closest to
the pulsed wall during rapid heating in three different pore sizes
were investigated. Afterward, the range in which multiple phases
existed within a pore before the pore was assumed isothermal was
investigated to determine where the model was beneficial. It was
determined that the existence of multiple melting fronts seen in Fig. 1
existed when the pulse period was less than  0.0033, after which
the profile shifted to a slow heating/cooling cycle and a single melting
front existed. It was also determined that the model was useful up to a
pulse period of  0.0175, when the pulse period was long enough
that the pore was fully solid or liquid throughout the whole cycle.
Future work in this subject will involve reduction of assumptions
made within the model, the influence of different latent heat values,
and use of the model to predict the behavior of systems under rapid
transient conditions.

Appendix: Numerical Validation


The validation of the numerical results of the model was done by
comparing to the results produced by [25] based on the Rayleigh,
Prandtl, Stefan, and Darcy numbers. To produce the same results, the
pore diameter, porosity, and effective thermal conductivity values
used were dP  0.0135L,  0.8, and kf  103 1   kP ,
respectively. The ratio of the thermal capacities (specific heat
multiplied by density) of the foam and PCM was set to unity, and the
interstitial Nusselt number was kept at NuD;i  5.9. The Stephan
c T H T C 
number was set as Ste  p;l H
, the Rayleigh number as
6
Ra  10 , Pr  50, and Da  102 . In the model, the input
temperature difference was set to T m;avg  0.3, and T H and T C were
constrained by T m2 T m1   0.02T H T C . Running the same
grid independence study as the results setup produced a grid with 57
foam nodes and 24 pore nodes. The results of the model were
compared to the results from [25] and produced the expected results
(see Fig. A1).

Fig. A1 Comparison of current model to numerical results from [26].

Acknowledgments
The authors would like to acknowledge the continued support of
Rolls-Royce North American Technologies, Inc. as well as any
individual involved therein.

References
[1] Hall, C. III, Glakpe, E., and Cannon, J., Modeling Cyclic Phase Change
and Energy Storage in Solar Heat Receivers, Journal of Thermophysics
and Heat Transfer, Vol. 12, No. 3, 1998, pp. 406413.
doi:10.2514/2.6352
[2] Fukai, J., , Hamada, Y., Morozumi, Y., and Miyatake, O., Improvement
of Thermal Characteristics of Latent Heat Thermal Energy Storage
Units Using Carbon-Fiber Brushes: Experiments and Modeling,
International Journal of Heat and Mass Transfer, Vol. 46, No. 23, 2003,
pp. 45134525.
doi:10.1016/S0017-9310(03)00290-4
[3] Fok, S. C., Shen, W., and Tan, F. L., Cooling of Portable Hand-Held
Electronic Devices Using Phase Change Materials in Finned Heat
Sinks, International Journal of Thermal Sciences, Vol. 49, No. 1, 2010,
pp. 109117.
doi:10.1016/j.ijthermalsci.2009.06.011
[4] Mesalhy, O., Lafdi, K., and Elgafy, A., Carbon Foam Matrices
Saturated with PCM for Thermal Protection Purposes, Carbon, Vol. 44,
No. 10, 2006, pp. 20802088.
doi:10.1016/j.carbon.2005.12.019
[5] Zhao, C. Y., Lu, W., and Tian, Y., Heat Transfer Enhancement for
Thermal Energy Storage Using Metal Foams Embedded with Phase
Change Materials (PCMs), Solar Energy, Vol. 84, No. 8, 2010,
pp. 14021412.
doi:10.1016/j.solener.2010.04.022
[6] Xiao, X., Zhang, P., and Li, M., Preparation and Thermal Characterization of Paraffin/Metal Foam Composite Phase Change Material,
Applied Energy, Vol. 112, Dec. 2013, pp. 13571366.
doi:10.1016/j.apenergy.2013.04.050
[7] Qu, Z. G., Li, W. Q., and Tao, W. Q., Numerical Model of the Passive
Thermal Management System for High-Power Lithium Ion Battery by
Using Porous Metal Foam Saturated with Phase Change Material,
International Journal of Hydrogen Energy, Vol. 39, No. 8, 2014,
pp. 39043913.
doi:10.1016/j.ijhydene.2013.12.136
[8] Baby, R., and Balaji, C., Experimental Investigations on Thermal
Performance Enhancement and Effect of Orientation on Porous Matrix
Filled PCM Based Heat Sink, International Communications in Heat
Mass Transfer, Vol. 46, Aug. 2013, pp. 2730.
doi:10.1016/j.icheatmasstransfer.2013.05.018
[9] Wierschke, K., Franke, M. E., Watts, R., and Ponnappan, R., Heat
Dissipation with Pitch-Based Carbon Foams and Phase-Change
Materials, Journal of Thermophysics and Heat Transfer, Vol. 20, No. 4,
2006, pp. 865870.
doi:10.2514/1.18463
[10] Kota, K., Chow, L., Du, J., Kapat, J., Leland, Q., and Harris, R.,
Numerical Analysis of Heat Storage Phenomenon in a Dual Latent
Heat Sink, Journal of Thermophysics and Heat Transfer, Vol. 23,
No. 1, 2009, pp. 148156.
doi:10.2514/1.36386
[11] Nayak, K. C., Saha, S. K., Srinivasan, K., and Dutta, P., A Numerical
Model for Heat Sinks with Phase Change Materials and Thermal
Conductivity Enhancers, International Journal of Heat Mass Transfer,
Vol. 49, Nos. 1112, 2006, pp. 18331844.
doi:10.1016/j.ijheatmasstransfer.2005.10.039
[12] Harris, K. T., Haji-Sheikh, A., and Agwu Nnanna, A. G., PhaseChange Phenomena in Porous MediaA Non-Local Thermal Equilibrium Model, International Journal of Heat and Mass Transfer, Vol. 44,
No. 8, 2001, pp. 16191625.
doi:10.1016/S0017-9310(00)00191-5
[13] Beckermann, C., and Viskanta, R., Natural Convection Solid/Liquid
Phase Change in Porous Media, International Journal of Heat and
Mass Transfer, Vol. 31, No. 1, 1988, pp. 3546.
doi:10.1016/0017-9310(88)90220-7
[14] Kaviany, M., Principles of Heat Mass Transfer in Porous Media,
Mechanical Engineering Series, Springer, Berlin, 1995, pp. 401404.
[15] Minkowycz, J., Haji-Sheikh, A., and Vafai, K., On Departure from
Local Thermal Equilibrium in Porous Media Due to a Rapidly Changing
Heat Source: The Sparrow Number, International Journal of Heat and
Mass Transfer, Vol. 42, No. 18, 1999, pp. 33733385.
doi:10.1016/S0017-9310(99)00043-5

Downloaded by PURDUE UNIVERSITY on May 26, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.T4523

JACKSON ET AL.

[16] Tian, Y., and Zhao, C.Y., A Numerical Investigation of Heat Transfer in
Phase Change Materials (PCMs) Embedded in Porous Metals, Energy,
Vol. 36, No. 9, 2011, pp. 55395546.
doi:10.1016/j.energy.2011.07.019
[17] Krishnan, S., Murthy, J. Y., and Garimella, S. V., A Two-Temperature
Model for the Analysis of Passive Thermal Control Systems, Journal of
Heat Transfer, Vol. 126, No. 4, 2004, pp. 628637.
doi:10.1115/1.1773194
[18] Nield, D. A., Notes on Convection in a Porous Medium: (i) An
Effective Rayleigh Number for an Anisotropic Layer, (ii) the Malkus
Hypothesis and Wavenumber Selection, Transport in Porous Media,
Vol. 27, No. 2, 1997, pp. 135142.
doi:10.1023/A:1006549930540
[19] Voller, V., and Cross, M., Accurate Solutions of Moving Boundary
Problems Using the Enthalpy Method, International Journal of Heat
Transfer, Vol. 24, No. 3, 1981, pp. 545556.
doi:10.1016/0017-9310(81)90062-4
[20] Bo, H., Viktoria, M., and Fredrik, S., Phase Transition Temperature
Ranges and Storage Density of Paraffin Wax Phase Change Materials,
Energy, Vol. 29, No. 3, 2004, pp. 17851804.
doi:10.1016/j.energy.2004.03.002
[21] Li, W. Q., Qu, Z. G., He, Y. L., and Tao, W. Q., Experimental
and Numerical Studies on Melting Phase Change Heat Transfer in
Open-Cell Metallic Foams Filled with Paraffin, Applied Thermal
Engineering, Vol. 37, May 2012, pp. 19.
doi:10.1016/j.applthermaleng.2011.11.001

819

[22] Regin, A. F., Solanki, S. C., and Saini, J. S., Latent Heat Thermal
Energy Storage Using Cylindrical Capsule: Numerical and Experimental Investigations, Renewable Energy, Vol. 31, No. 13, 2006,
pp. 20252041.
doi:10.1016/j.renene.2005.10.011
[23] Boomsma, K., and Poulikakos, D., On the Effective Thermal
Conductivity of a Three-Dimensionally Structured Fluid-Saturated
Metal Foam, International Journal of Heat and Mass Transfer, Vol. 44,
No. 4, 2001, pp. 827836.
doi:10.1016/S0017-9310(00)00123-X
[24] Fourie, J. G., and Du Plessis, J. P., Pressure Drop Modelling in Cellular
Metallic Foams, Chemical Engineering Science, Vol. 57, No. 14, 2002,
pp. 27812789.
doi:10.1016/S0009-2509(02)00166-5
[25] MATLAB and Simulink, Software Package, Ver. 2014a, MathWorks,
Natick, MA, 2014.
[26] Krishnan, S., Murthy, J. Y., and Garimella, S. V., Analysis of SolidLiquid Phase Change Under Pulsed Heating, Journal of Heat Transfer,
Vol. 129, No. 3, 2007, pp. 395400.
doi:10.1115/1.2430728
[27] Eftekhar, J., Haji-Sheikh, A., and Lou, D. Y. S., Heat Transfer
Enhancement in a Paraffin Wax Thermal Storage System, Journal of
Solar Energy Engineering, Vol. 106, No. 3, Aug. 1984, pp. 299306.
doi:10.1115/1.3267599

Вам также может понравиться