You are on page 1of 26

Accepted Manuscript

Numerical simulation of effective thermal conductivity and pore-scale melting


process of PCMs in foam metals

Gang Wang, Gaosheng Wei, Chao Xu, Xing Ju, Yanping Yang, Xiaoze Du

PII: S1359-4311(18)31765-4
DOI: https://doi.org/10.1016/j.applthermaleng.2018.10.106
Reference: ATE 12850

To appear in: Applied Thermal Engineering

Received Date: 23 March 2018


Revised Date: 18 October 2018
Accepted Date: 23 October 2018

Please cite this article as: G. Wang, G. Wei, C. Xu, X. Ju, Y. Yang, X. Du, Numerical simulation of effective thermal
conductivity and pore-scale melting process of PCMs in foam metals, Applied Thermal Engineering (2018), doi:
https://doi.org/10.1016/j.applthermaleng.2018.10.106

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Numerical simulation of effective thermal conductivity and pore-scale

melting process of PCMs in foam metals


Gang Wang, Gaosheng Wei, Chao Xu, Xing Ju, Yanping Yang, Xiaoze Du
School of Energy, Power and Mechanical Engineering, Key Laboratory of Condition Monitoring and Control for
Power Plant Equipment of Ministry of Education, North China Electric Power University, Beijing 102206, China.

Abstract:
Different foam metals combined with paraffin and other materials were analyzed to determine their effective

thermal conductivity and the macroscopic thermophysical properties of the composite materials. A W-P model

composed of six tetrakaidecahedrons and two irregular dodecahedrons was used to simulate the melting heat transfer

process in open foam metal at pore-scale under constant temperature. The results show that the porosity and

conductivity of the foam metal and the conductivity of the phase change material (PCM) have a significant influence

on the effective thermal conductivity of the composite PCM, while the pore size has no obvious influence. The

effective thermal conductivity of composite PCMs increased with increasing foam metal thermal conductivity, and

increased more rapidly with lower foam metal porosity. The effective thermal conductivity of composite PCMs is

related to the ratio of foam metal conductivity to PCM conductivity. The microstructure of the foam metal had an

obvious effect on the solid-liquid phase distribution during the PCM melting process, where the heat was transferred

mainly through the melted liquid PCM field. Conduction was the dominant heat transfer mechanism, and natural

convection in the liquid PCM was weak for the confinement of foam metals. For heat transfer during the PCM

melting process, conduction through the skeleton of the porous metal played the most important role. The PCM

adjacent to the heating source and foam metal frame melted first, with the fusion zone gradually spreading to the

pore center. The melting rate of the PCM increased with increasing boundary temperature and thermal conductivity

of the foam metal, but decreased as foam metal porosity increased. During the melting process, the liquid phase

fraction did not linearly grow with time; the melting rate was very large at the initial stage, but decreased gradually

with time.
Keywords: phase change material (PCM); heat transfer enhancement; metal foam; effective thermal conductivity;

phase change process;

1. Introduction

Adding an energy storage system to a renewable energy harvesting system is one method used to balance and

adjust the energy output in order to stabilize the energy output of the system. One option is thermal energy storage

(TES) systems, which have a low construction cost, large energy storage density, and the ability to store energy

directly. Phase change materials (PCM) possess a large heat capacity, and when used in a TES system the thermal

energy can be stored at nearly constant temperature during the phase change process. Because of these features,

phase change TES systems have a potentially broad application for use in solar heating systems and solar energy

generation systems, as well as use in large-scale industry [1]. However, most PCMs have a low thermal conductivity,

and so enhancing thermal conductivity is a very important subject for phase change TES technology. Cárdenas [2]

and Wei [3] have written comprehensive reviews on the study of thermal conductivity enhancement of PCMs for a

variety of strengthening heat transfer methods, summarizing the work of many other researchers who have

completed both experimental and theoretical studies on improving the thermal conductivity of various PCMs.

Foam metals have many excellent characteristics that make them suitable for use in a TES, such as high

porosity, high specific surface area, and high thermal conductivity. In addition, foam metals can be used as the

matrix of a PCM to enhance its thermal conductivity while having a negligible effect on the heat storage capacity of

the PCM [4-6]. The arbitrarily distributed struts inside a foam metal can also strengthen the heat transfer process.

Zhao et al. [7] and Zhou et al. [8] found that foam metal was more effective than expanded graphite for enhancing

heat transfer performance. The experimental results of Tian and Zhang [9-11] also showed that a porous metal foam

could effectively increase the heat transfer rate of an energy storage system. Historically, adding foam metal into a

PCM has been a popular way to increase thermal conductivity of the composite material. Among different foam

metals, foam copper, foam aluminum, and foam nickel have been the three most commonly studied materials for

enhancing the thermal conductivity of a PCM. Zhang et al. [12], Zhang et al. [13], Xiao [14], Vadwala [15], and

Wang [16] have all measured the effective thermal conductivity of a foam copper/paraffin composite. Sung-Tae

Hong [17] measured the effective thermal conductivity of three paraffin/foam aluminum composites under different

temperature conditions using the steady-state method. Xu [18] prepared a paraffin/foam nickel specimen and

measured its effective thermal conductivity using the transient plane method.
Researches on the effective thermal conductivity of porous media can be traced back to the 19th century. The

methods for measuring the effective thermal conductivity including experimental method, numerical simulation

method and empirical correlation method. But both experimental method and numerical method are time-consuming

and subject to various factors. Various empirical correlations based on experimental data or theoretical analysis can

predict the effective thermal conductivity directly, accurately and reasonably. Boomsma and Poulikakos[19]

established a three-dimensional positive tetrahedral skeleton model based on the minimum surface theory to predict

the effective thermal conductivity of foam metals. Dul 'nev [20] established the cubic lattice model and predicted the

effective thermal conductivity of the foam metal by using the effective thermal resistance method. Singh and

Kasana[21] derived a new empirical formula to predict the effective thermal conductivity of foam metal based on

the non-linear heat flow vector caused by different geometric shapes of porous media and different thermal

conductivity of continuous phase. Bhattacharya[22] predicted the effective thermal conductivity of the foam metal

through theoretical analysis and experimental study. Mendes[23] proposed a new prediction model, which includes

two adjustable parameters.

While the effective thermal conductivity cannot reveal the phase change process, it can help characterize the

macroscopic heat transfer phenomenon. Many researchers have conducted relevant studies on the phase change

process of PCMs in a foam metal. In numerous experimental studies, Zhang et al. [24] and Thapa et al. [25]

observed the heat transfer of paraffin in foam copper. The experimental results by Zhuo et al. [26-29] showed that

the phase change rate could be reduced by 28.1 times for solid state PCMs and by 3.1 times for liquid state PCMs

using foam copper. Hong et al. [4] and Chen et al. [5, 6] investigated the heat transfer of paraffin in open-cell foam

aluminum. The experimental results of Allen and Xie [30, 31] showed that heat transfer performance could be

increased 25.2% by foam aluminum. Tian et al. [32] created simulations of heat transfer of a PCM in porous metals

based on the two-equation non-equilibrium heat transfer model and numerically analyzed the results. Li [33]

conducted a numerical investigation of thermal behavior of sodium nitrate (NaNO3) inside a porous metal matrix by

using an ideal cube-shaped unit structure. Hu et al. [34] investigated the heat transfer process of a paraffin PCM in

aluminum foam metal by constructing a body-centered-cubic (BCC) lattice unit cell model embedded with spherical

micro-pores. Sundarram et al. [35] constructed a three-dimensional face-centered-cubic (FCC) structure model to

investigate the pore size effect on the phase change process of a PCM in aluminum foam metal. Chen [6] conducted

a pore-scale numerical study of the melting of paraffin in aluminum foam metal in a two-dimensional computational

domain using the Lattice-Boltzmann method.

The heat transfer process involves conduction, convention, and radiation, but in the small pores of foam metal,
it is difficult to represent every process. Although there have been many experimental studies and theoretical

analyses conducted on the solid-liquid phase change heat transfer process of PCMs in foam metals, there have been

few measurements or observations of the heat transfer process at the pore scale. In this paper, we convert a number

of different heat transfer models into a considerable heat conduction problem in a macroscopic way to represent the

heat transfer mechanism of porous media, such that the effective thermal conductivity represents the material’s heat

transfer ability. We chose to use the three most common foam metals (copper, aluminum, nickel) as a matrix to

simulate the effective thermal conductivity derived using the W-P model. The solid-liquid phase change heat transfer

process of the PCMs in various porous foam metals are simulated at pore-scale and combined with analysis of the

effective thermal conductivity so that the effects of boundary conditions and the porosity of the foam metals on the

heat transfer process can be analyzed. Finally, we characterize the heat transfer ability of the various composites by

measuring the effective thermal conductivity to quantitatively judge the heat transfer reinforcement of PCMs by

using foam metals.

2. Theoretical model

2.1 Geometric model construction

The most commonly used technique to manufacture foam metals is to blow a specific type of foaming gas into

the molten metal so that the gas bubbles move freely in the molten metal and a balanced state is reached between the

liquid metal and the gas bubbles. This is called the minimum surface energy state. The surface energy minimization

theory is the first hypothesis for construction of a foam metal geometric model [36]. Weaire and Phelan [37]

discovered that the optimal minimal surface energy at a given volume formed a single cell structure composed of six

tetrakaidecahedrons and two irregular dodecahedrons

For this work, we used the Surface Evolver software and the Weaire and Phelan optimal minimal surface

energy theory to develop an element structure referred to as the W-P model. Fig. 1(a) illustrates the crystal cell

structure, and shows that the surface of the cell is a triangle, while Fig. 1(b) shows the frame structure. It is assumed

that the cross section of the skeleton is circular, and that the skeleton is connected together through a uni-diameter

ball. Although this structure is not identical to the real frame of a foam metal, it can be correctly used to analyze heat

transfer for a medium with high porosity [38]. The porosity of a foam metal can be changed by designing the

cross-sectional diameter of the frame and array spreading of the skeleton in the x/y/z directions to fill the entire

space. The frame structure of the foam metal is shown in Fig. 2(a). The computational domain can be extracted from
this frame structure, as shown in Fig. 2(b).

2.2. Mathematical model

In order to simplify the numerical calculations for the simulation, we made the following two assumptions

when using the above model to solve for effective thermal conductivity:

(1) Assuming the fluid is still and ignoring the effect of radiation, the energy equations for the solid and fluid

phases in the steady state heat transfer process are:

(s Ts )  0
(1)

(2)
( f T f )  0

where λ is the thermal conductivity and T is temperature. Subscript s denotes a solid and f denotes a fluid.

(2) Assume that the two-phase interface is a coupled boundary condition (i.e. at thermal equilibrium), and heat

flow through the interface is driven by the temperature gradient of the two phases. Then the two-phase boundary

satisfies
(3)
Ts  T f
(4)
sTs n   f Tf n

and the heat flow in a given direction can be calculated by


  s Ts d  s    T d 
f f f
s f (5)
q 

where   s   f is the cross-sectional area perpendicular to the direction of heat flow  , and Thot and Tcold are
the boundary temperatures on two sides. The heat flow through a given direction can be simplified to

Thot  Tcold
q  eff (6)
L
where L is the length along the direction of the heat flow  .

And the thermal effective conductivity can be calculated by

  q  dA
eff = A (7)
T  A

where A is the area along the direction of the heat flow  .

The following assumptions were made when using the model to simulate the phase change heat transfer

process:

(1) The thermophysical properties of the PCMs are constant in the phase change temperature range.
(2) The volume of paraffin does not change in the phase change process.

(3) Liquid paraffin is an incompressible Newtonian fluid.

(4) Natural convection obeys the Boussinesq hypothesis.

(5) Liquid paraffin experiences a laminar flow.

Based on the above assumptions, the governing equations for a liquid paraffin PCM are
(7)
u  0
(8)
u
f   f (u  )u  P   f  2 u   f g (T f  Tm )  Au
t
(9)
C (1  f1 ) 2
A
S  f13

where, f l is the liquid fraction, C  0.001 , and S  0.000001 . The value of f l is 0 and 1 when the PCM is

solid and liquid, respectively, and ranges between 0 and 1 in the indeterminate region.

The energy equation of the PCM can be expressed as

T f f l (10)
 f c pf   f c pf u  T f    ( λ f T f )   f L
t t

where
(11)
f l  0, T  Tsolidus
f l  1, T  Tsolidus
T T solidus
fl  , Tsolidus  T  Tliquidus
Tliquidus  Tsolidus

In order to ensure the stability of the calculation, it is assumed that the temperature of the phase change process

occurs in a small range Tsolidus  T  Tliquidus, such that Tsolidus is 0.1 °C below the melting temperature and Tliquidus

is 0.1 °C higher than the melting temperature.

Finally, the energy equation of the foam metal skeleton can be written as
Ts
 s c ps    ( λs Ts ) (12)
t

2.3. Mesh generation and mesh independence verification

The finite volume software ANSYS FLUENT 15.0 was used to solve the mathematical models developed in

this work, with the assumption that the PCM fully fills the cell depicted in Fig. 2(b). The geometry was created in

CAD and the mesh generated in ICEM. The computed field was 5.08×5.08×5.08 mm, divided into a fluid area and a

solid area. The mesh was generated using the octree method.
The effective thermal conductivity was calculated at steady state, and a second order scheme used for

discretizing the solid metal skeleton. When the energy equations converged, the heat flux through the cross section

of the representative elementary volume could be integrated by the heat flux of the hot and cold sections (the heat

flux in the hot and cold sections are the same when the equations converged). The effective thermal conductivity

was then calculated by the Fourier law in Equation (6) with a known heat flow and temperature gradient. The

boundary conditions used were Thot  298.15K and Tcold  293.15K . The effective thermal conductivity of the

different foam metal/PCM composites were calculated using this method to judge the heat transfer reinforcement of

the PCM by the different foam metals.

Four meshes were generated to verify independence from mesh size, with total mesh sizes of 234453, 1089682,

1791757, and 3608870, respectively. The calculated thermal conductivities of the foam metal/air composites and

foam metal/water composites are shown in Fig. 3(a). The results using the latter three meshes are almost the same,

and only the result of the first mesh is obviously different from the others. Based on these results, the third mesh size

was selected to simulate the effective thermal conductivity of the composite PCMs.

The solidification/melting model was adopted for solving the melting process of each composite PCM, and a

prior second order scheme used to discretize the convection equation. The SIMPLE algorithm was used to solve the

equations. The convergence factors for the momentum equation, pressure correction equation, energy equation, and

liquid fraction equation were 0.7, 0.3, 1.0, and 0.6 respectively. The residual error convergence criterion was 10-8 for

the energy equation and 10-5 for the rest of the equations, and the time step was dt=0.01 s. The meshes used in the

foam metal skeleton region and the phase change material region were unstructured meshes. Two separate mesh

projects of 461997 and 1089682 unstructured meshes were designed for these regions. Fig. 3(b) shows that the

results of these two mesh projects in both the melting process and liquid fraction regions are almost the same,

verifying the independence of mesh size. Thus, a mesh size of 461997 was chosen to simulate the melting process of

the PCM.

Paraffin was selected as the PCM for the numerical calculations, and its major properties are shown in Table 1.

The kinematic viscosity was μ = 2.63×10-2 - 6.87×10-5 T. Note that we assumed the thermophysical properties of

paraffin are constant and that the volume did not change during the phase change process. Liquid paraffin is an

incompressible Newtonian fluid that obeys Boussinesq’s hypothesis of natural convection flow, which assumes that

the phase change process occurs in the narrow temperature range Tsolidus = Tm  0.1 °C and Tliquidus  Tm  0.1 °C, where

Tm is the phase change temperature. The sensible heat was also considered in the numerical heat transfer analysis of
the melting process, even though its value is very small compared with the latent heat. The initial temperature was

300 K. The constant temperature boundary condition at the lower surface of the domain was T = 350 K starting from

time t=0, while the upper boundary of the domain was an adiabatic condition and the side faces were symmetric

boundary conditions. The numerical calculation terminated when all of the phase change material was converted to

liquid phase, so any overheating effect on the energy storage system was ignored.

3. Results and discussion

3.1 Effective thermal conductivity of the composite

Conduction in the foam metal/filling material composite can be divided into two parts, with some heat

conduction occurring through the metal skeleton and the remainder through the filling material. The factors affecting

the conductivity of the composites can thus be analyzed by separately considering the foam metal thermal properties

and the filling material thermal properties.

3.1.1. Effect of foam metal pore size on effective thermal conductivity

Foam aluminum 6101 was chosen as the simulation object, with a porosity of 86.42% and pore sizes of 10 PPI,

20 PPI, 30 PPI, and 40 PPI. Air and water were chosen as additives. Fig. 4 shows the effect of pore size of the foam

metal on the effective thermal conductivity. For the controlled porosities, the conductivities of the different pore

sizes in the foam aluminum/air composites are 13.167 W/(m·K), 13.167 W/(m·K), 13.168 W/(m·K) and

13.168 W/(m·K) for pore sizes of 10 PPI, 20 PPI, 30 PPI, and 40 PPI, respectively. The conductivities of the same

pore sizes in foam aluminum/water composites are 13.802 W/(m·K), 13.802 W/(m·K), 13.801 W/(m·K) and

13.803 W/(m·K). These results agree with the results calculated by using the regular cube unit cell model of Ashby

[34], which shows that the pore size has no obvious effect on thermal conductivity. From the simulation results, we

can conclude that the effective thermal conductivity has no obvious relationship to the pore size of foam metals.

3.1.2. Effect of foam metal porosity on effective thermal conductivity

Different porosities of foam aluminum/air composites, foam aluminum/water composites, and foam

copper/paraffin composites were simulated such that all of the foam metals had a uniform pore size of 10 PPI with a

heat source of 298 K. The calculated results are compared to data from the literature in Fig. 5. The simulated results

from this study and other predicted results (lines) and experimental results (points) from the literature are quite

consistent, and there are almost no differences among the results when the porosity is greater than 0.93. In the foam

aluminum/air composites, the conductivity decreases from 19.566 to 1.902 as the porosity of the foam aluminum
increases from 0.812 to 0.975. Similarly, in the foam aluminum/water composites and foam copper/paraffin

composites the conductivities decreases from 20.211 and 24.364 to 2.501 and 0.659, respectively, for the same

change in porosity. The effective thermal conductivity displays a rapid almost linear decline with increasing porosity

of foam metal in all of the composites. The reason for this result could be the large difference in conductivities of the

foam metal and the filling materials. As the porosity increases there is less of the foam metal in the composite and

more of the filling material, which means the effective thermal conductivity should decrease as the filling materials

all have a lower thermal conductivity than the foam metal. The simulated results from this work are most similar to

the predicted results of Dul’nev [20], because the model used by Dul’nev is only related to porosity while the

models presented by others involve other parameters in addition to porosity. Note that the simulation results are a

little higher than the experimental results depicted in Fig. 5. Possible reasons for this discrepancy include heat loss,

thermal contract resistance, and simplification of the real model. All of these factors will cause the simulation to be

too ideal and to ignore some small factors.

3.1.3 Effect of component conductivity on effective thermal conductivity

In this section, we first analyze the effect of the thermal conductivity of the filling material on effective thermal

conductivity with the thermal conductivity of the foam metal remaining unchanged. The foam copper/paraffin

composite was selected as the sample, and the conductivity of the paraffin artificially set from 1 to 398 W/(m·K). In

Fig. 6, the ratio of composite conductivity λ to copper conductivity λs is the y-coordinate and the ratio of filling

material conductivity λf to copper conductivity λs is the x-coordinate. It can be seen in the figure that the influence of

λf on λ is very small when λf is much less than λs. Conversely, λf has a large influence on λ when λf is close to λs and

λ will approach λs. In conclusion, the effective thermal conductivity of the composite is closely related to the ratio of

foam metal conductivity to filling material conductivity.

Next, the effect of the foam metal conductivity on the effective thermal conductivity was analyzed by keeping

the thermal conductivity of the filling material unchanged. Paraffin was chosen as the filling material, and Ni, Al,

and Cu chosen for the foam metal matrix. For this analysis, 18 results using three different composites under six

porosity conditions were calculated, with the results shown in Fig. 7. The results show that the Cu matrix composite

has the largest effective thermal conductivity for a given porosity, while the Ni matrix composite has the smallest

conductivity. These respective rankings occur irrespective of porosity condition. The conductance relationship of the

three metals mentioned above is the same with the three composites. This suggests that the effective thermal

conductivity of the composite is mainly dependent on the thermal conductivity of the foam metal. That is, the
conductivity of the composite will increase with the conductivity of the component foam metal. From Fig. 7 we can

also see that a lower porosity of foam metal has a greater influence on the foam metal conductivity and on the

effective thermal conductivity. This is because the metal content in the composites gradually increases as porosity

decreases. Thus, the influence of the foam metal conductivity on the effective conductivity obviously increases.

3.2 The melting process of foam metal/paraffin composites

3.2.1 Phase and temperature distribution in the melting process

Figure 8 gives the solid-liquid phase and temperature distribution at different locations when the liquid fraction

is 0.25 during the melting process of a foam aluminum/paraffin composite. The red part in the figure denotes the

melted paraffin, while and the blue part denotes the foam metal and the solid paraffin. From the figure, we can see

that the melting process starts from the lower heat source surface, and the heat transmits upward. This agrees with

the results of Feng [49], who used a volume-averaged model. Fig. 8(b) illustrates how the temperature rise of the

foam metal frame is apparently faster than that of the paraffin due to its higher thermal conductivity, despite the

foam metal geometry being different at different cross sections. The figure also shows that the paraffin surrounding

the foam metal frame is the first to reach the phase change temperature and melt. The same phenomenon was

observed by infrared camera in an experiment conducted by Chen [50]. This result indicates that the phase change

diagram is greatly affected by the distribution of the foam metal matrix—that is, the melted paraffin surrounding the

foam metal frame—and that conduction is the dominant heat transfer mode in the phase change process.

Figure 9 describes the solid-liquid phase and change in temperature distribution with time at the cross section

of Z/L=0.5. L is the total height of the computational domain, Z is the height of the section selected in the L direction,

and Z/L is the relative position of the section selected in the L direction. The figure shows that the paraffin at the

lower heat source surface is the first to melt, followed by the paraffin around the foam metal frame from bottom to

top and then gradually spreading to the center of the pores. Therefore, the solid-liquid interface of the PCM extends

from the bottom to the top of the domain, and from the interface of the paraffin and foam metal frame to the center

of each pore. Note that for some levels, heat conduction has increased the foam metal frame temperature above the

phase change temperature of the PCM even though the PCM at that layer is still not fully melted. Similar

phenomena were analyzed by Deng [51] and Yang [52], who found that the temperature gradient in the liquid phase

region was larger than the temperature gradient in the solid phase region. This indicates that the heat coming from

the heating surface and the foam solid frame is mainly absorbed by the liquid PCM instead of the solid PCM. The

biggest temperature gradient is closer to the melting phase interface, which indicates that most of the heat has been
transformed into latent heat in the PCM. Note that the solid-liquid distribution and the temperature distribution in

the whole simulated domain was not fully identical, and this is mainly due to the irregular distribution of the foam

metal frame structure.

3.2.2 Change in temperature during the melting process

Figure 10 gives the change in temperature for different monitoring points in the domain. Point 2 in Fig. 10 is

located on the foam metal frame, while points 1 and 3 are located at different points in the PCM at the central axis of

the domain. The temperature of each monitor point in the PCM shown in the figure first increases to a melting

temperature of 317 K, then remains constant until the solid-liquid phase interface reaches the monitor point. This

result agrees well with the experimental results of Zhang [53]. While the temperature of the monitor point on the

foam metal frame rapidly increases at the initial stage, the rate of change slows as the phase change occurs for the

PCM around the frame. The temperature of the melted PCM increases with time, creating a temperature difference

between the melted PCM and the melting interface. When the PCM is fully melted, the temperature of the PCM

rapidly increases until reaching an equilibrium temperature of 350 K because constraints imposed by the solid

paraffin layer have been removed. The observed changes in temperature are consistent with the pure sensible heat

storage process.

Figure 11 shows the change in heat flux at the bottom surface for different heat source temperatures and

different foam metal materials. The heat flux at the bottom surface of a foam copper/paraffin composite is illustrated

in Fig. 11(a). The three scatter lines represent the change in heat flux with time under three different heat source

temperature conditions. At the initial moment, the temperature difference between the heat source and the initial

temperature is its maximum and the heat flux is largest, but this difference decreases with time. This phenomenon

appears in all three heat source temperature conditions. Because the total heat storage is certain, a higher heat source

temperature and resulting larger heat flux results in a shorter heat storage time. The heat flux at the bottom surface

of three different composites are shown in Fig. 11(b) for a heat source temperature of 350 K. Similar to the results of

Fig. 11(a), the heat flux is very large at the start. The heat flux of the foam copper/paraffin composite is largest

during the entire process while the heat flux of the foam nickel/paraffin composite is smallest. This is because the

foam copper/paraffin composite has the largest effective thermal conductivity. Due to Fourier’s Law, the heat flux is

proportional to thermal conductivity, so the heat flux of the foam copper/paraffin is highest and the heat flux of the

foam nickel/paraffin is lowest in Fig. 11(b).

A decrease in heat flux over time occurs for all three composites. The reduction of heat flux in both Fig. 11(a)
and Fig. 11(b) occurs due to the combination of two factors, the temperature difference and the melting process of

the paraffin. The temperature difference is largest at the initial moment and then decreases with time. Since the heat

flux is inversely proportional to the temperature difference, the heat flux will also decrease with time. As the

temperature of the paraffin approaches 317 K, the paraffin melts and a liquid film forms between the heating surface

and the solid phase, and the thickness of this film increases with time. This growing liquid film leads to an increase

in thermal resistance between the heating surface and the solid PCM, which apparently decreases heat flux. The

combined effect of the above two factors leads to the slowly reduced decreasing rate of heat flux.

3.2.3 Change in liquid fraction during the melting process

As with the previous section, three composites were simulated using a heating source temperature of 370 K,

while a foam aluminum/paraffin composite was simulated under three heating source temperature conditions of 330

K, 350 K, and 370 K.

Figure 12 gives the change in liquid fraction with different heating boundary temperatures, different foam

metal materials, and different porosity of foam metals. In Fig. 12(a), the liquid fraction is obviously affected by the

use of different foam metal materials because of the quantitative conductivity relation. For any given moment, the

liquid fraction of the foam copper/paraffin composite is largest and the liquid fraction of the foam nickel/paraffin

composite is smallest. This is the same relationship as with the effective thermal conductivity. Thus, the conductivity

of the foam metal has an obvious effect on the melting rate. In Fig. 12(b), the liquid phase fraction does not grow

linearly with time. The melting rate is large at the initial stage, and decreases gradually with time. The heat transfer

occurs mainly through the melted liquid PCM, so the melting rate decreases because the thickness of the liquid PCM

increases with time. This agrees well with the experimental results of Zhang [54], and both results verify that

conduction is the major pathway of heat transfer inside a foam metal. Furthermore, Fig. 12(c) shows that the melting

rate increases with decreasing porosity. Thus, the foam metal matrix structure can enhance the heat transfer of a

PCM.

4. Conclusions

Using the pore-scale representative elementary volume (REV) method and finite element software, the effective

thermal conductivity and the melting heat transfer process of PCMs in porous foam metals were numerically

simulated in this study. The effective thermal conductivity of different foam metals combined with paraffin and

other materials was analyzed to determine the macroscopic thermophysical properties of the composite, although it
could not reveal the melting process of the PCM at pore-scale. The melting process of PCMs in a foam metal was

numerically analyzed using different boundary conditions and the effective thermal conductivity and pore-scale

melting process analyzed. The results can be concluded as follows:

(1) The effective thermal conductivity was inversely proportional to the porosity of the foam metal and had no

significant relationship to pore size at ambient temperature. At high porosity ranges (ε>0.9), the simulation results

were in good agreement with other experimental and predicted results found in the literature.

(2) The effective thermal conductivity of the composite PCMs increased with increased foam metal thermal

conductivity, and increased more rapidly when the porosity of the foam metal was much lower. The effective

thermal conductivity of the composite PCM is related to the ratio of the foam metal conductivity to PCM

conductivity.

(3) The PCM adjacent to the heating source and foam metal frame was the first to melt, and melting gradually

spread to the center of the pores. The temperature rise of foam metal frame was faster than the increase in

temperature of the PCM.

(4) The temperature gradient in the liquid phase region was larger than the temperature gradient in the solid phase

region. The largest temperature gradient occurred closest to the melting phase interface. This indicates that the heat

coming from the heating surface and the foam solid frame was mainly converted to latent heat in the PCM.

(5) The melting rate of the PCM increased with rise of boundary temperature and thermal conductivity of the

foam metal, and increased as the porosity of the foam metal decreased.

(6) The liquid phase fraction did not grow linearly with time. The melting rate was very large at the initial stage,

and decreased gradually with time. Heat transfer occurred mainly through the melted liquid PCM, and conduction

was the dominant heat transfer model, with weak natural convection in the liquid PCM for confinement of the foam

metal.

Acknowledgement

This research was financially supported by the National Natural Science Foundation of China (No. 51776066),

the Fundamental Research Funds for the Central Universities (No. 2018ZD04) and the National Scholarship

Program of China (ICET2018).

Reference

[1] P. Giménez, A. Jové, C. Prieto, S. Fereres, Effect of an increased thermal contact resistance in a salt
PCM-graphite foam composite TES system, Renewable Energy. 106(2017) 321-334.

[2] B. Cárdenas, N. León, High temperature latent heat thermal energy storage: Phase change materials, design

considerations and performance enhancement techniques, Renewable and Sustainable Energy Reviews.

27(2013) 724-737.

[3] G.S. Wei, G. Wang, C. Xu, X. Ju, L.J. Xing, X.Z. Du, Y.P. Yang, Selection principles and thermophysical

properties of high temperature phase change materials for thermal energy storage: A review, Renewable and

Sustainable Energy Reviews. 2017.

[4] S.T. Hong, D.R. Herling, Open-cell aluminum foams filled with phase change materials as compact heat sinks,

Scr Mater. 55(2006) 887-890.

[5] Z.Q. Chen, M.W. Gu, D.H. Peng, Heat transfer performance analysis of a solar flat-plate collector with an

integrated metal foam porous structure filled with paraffin, Appl Therm Eng. 30(2010) 1967-1973.

[6] Z.Q. Chen, S. Juan, Experimental and numerical study on melting of phase change materials in metal foams at

pore scale, Int J Heat Mass Transf. 72(2014) 646-655.

[7] C.Y. Zhao, W. Lu, Y. Tian, Heat transfer enhancement for thermal energy storage using metal foams

embedded within phase change materials (PCMs), Sol Energ. 84(2010) 1402-1412.

[8] D. Zhou, C.Y. Zhao, Experimental investigations on heat transfer in phase change materials (PCMs) embedded

in porous materials, Appl Therm Eng. 31(2011) 970-977.

[9] Z. Acem, J. Lopez, E. Palomo, KNO3/NaNO3 graphite materials for thermal energy storage at high temperature:

Part I. Elaboration methods and thermal properties, Appl Therm Eng. 30(2010) 1580–1585.

[10] Y. Tian, C.Y. Zhao, A numerical investigation of heat transfer in phase change materials (PCMs) embedded in

porous metals, Energy. 36(2011) 5539–5546.

[11] C.B. Zhang, L.Y. Wu, Y.P. Chen, Study on solidification of phase change material in fractal porous metal

foam, Fractals-Complex Geom Patterns Scaling Nat Soc. 23(2015) 8.

[12] Z. Tao, J. Yu, H. Gao, Measurement of Thermal Parameters of Copper Foam/Paraffins Composite PCM Using

Transient Plane Source(TPS) Method, Acta Energiae Solaris Sinica. 31(2010) 604-609.

[13] X. Xiao, P. Zhang, M. Li, Effective thermal conductivity of open-cell metal foams impregnated with pure

paraffin for latent heat storage, International Journal of Thermal Sciences. 81(2014) 94-105.

[14] X. Xiao, P. Zhang, M. Li, Preparation and thermal characterization of paraffin/metal foam composite phase

change material, Applied Energy. 112(2013) 1357-1366.

[15] P. Vadwala, TheXirmal Energy Storage in Metal Foams filled with Paraffin Wax, School of Graduate Studies
– Theses. 2012.

[16] C. Wang, T. Lin, N. Li, H.P. Zheng, Heat transfer enhancement of phase change composite material: Copper

foam/paraffin, Renewable Energy. 96(2016) 960-965.

[17] S.T. Hong, D. R. Herling, Effects of Surface Area Density of Aluminum Foams on Thermal Conductivity of

Aluminum Foam‐ Phase Change Material Composites, Advanced Engineering Materials. 9(2010) 554-557.

[18] W.Q. Xu, X.G. Yuan, L. Zhen, Study on effective thermal conductivity of metal foam matrix composite phase

change materials, Journal of Functional Materials. 40(2009) 1329-1332+1337.

[19] K. Boomsma, D. Poulikakos, On the effective thermal conductivity of a three-dimensionally structured

fluid-saturated metal foam, International Journal of Heat & Mass Transfer. 44(2001) 827-836.

[20] G.N. Dul'Nev, Heat transfer through solid disperse systems, Journal of Engineering Physics. 9(1965) 275-279.

[21] R. Singh, H.S. Kasana, Computational aspects of effective thermal conductivity of highly porous metal foams,

Applied Thermal Engineering. 24(2004) 1841-1849.

[22] A. Bhattacharya, V.V. Calmidi, R.L. Mahajan, Thermophysical properties of high porosity metal foams,

International Journal of Heat & Mass Transfer. 45(2002) 1017-1031.


[23] M. Mendes, S. Ray, D. Trimis.An improved model for the effective thermal conductivity of open-cell porous
foams. Int. J. Heat Mass Tran, 75 (2014): 224-230.
[24] P. Zhang, Z.N. Meng, H. Zhu, Experimental and numerical study of heat transfer characteristics of a

paraffin/metal foam composite PCM, Energ Procedia. 75(2015) 3091-3097.

[25] S. Thapa, S. Chukwu, A. Khaliq, L. Weiss, Fabrication and analysis of small-scale thermal energy storage with

conductivity enhancement, Energ Convers Manag. 79(2014) 161-170.

[26] L. Zhuo, Z.G. Wu, Numerical study on the thermal behavior of phase change materials (PCMs) embedded in

porous metal matrix, Sol Energy. 99(2014) 172–184.

[27] S.S. Feng, Y. Zhang, M. Shi, T. Wen, T.J. Lu, Unidirectional freezing of phase change materials saturated in

open-cell metal foams, Appl Therm Eng. 88(2015) 315–321.

[28] S. Mancin, A. Diani, L. Doretti, K. Hooman, L. Rossetto, Experimental analysis of phase change phenomenon

of paraffin waxes embedded in copper foams, Int J Therm Sci. 90(2015) 79–89.

[29] J.L. Yang, L.J. Yang, C. Xu, X.Z. Du, Numerical analysis on thermal behavior of solidliquid phase change

within copper foam with varying porosity, Int J Heat Mass Transf. 84(2015) 1008–1018.

[30] M.J. Allen, T.L. Bergman, A. Faghri, N. Sharifi, Robust heat transfer enhancement during melting and

solidification of a phase change material using a combined heat pipe-metal foam or foil configuration, Heat
Transf-Trans Asme. 137(2015) 12.

[31] B. Xie, W.L. Cheng, Z.M. Xu, Studies on the effect of shape-stabilized PCM filled aluminum honeycomb

composite material on thermal control, Int J Heat Mass Transf. 91(2015) 135–143.

[32] Y. Tian, C.Y. Zhao, A numerical investigation of heat transfer in phase change materials (PCMs) embedded in

porous metals, Energ. 36(2011) 5539-5546.

[33] Z. Li, Z.G. Wu, Numerical study on the thermal behavior of phase change materials (PCMs) embedded in

porous metal matrix, Sol Energ. 99(2014) 172-184.

[34] X. Hu, S.S. Patnaik, Modelling phase change material in micro-foam under constant temperature condition, Int

J Heat Mass Transf. 68(2014) 677-682.

[35] S.S. Sundarram, W. Li, The effect of pore size and porosity on thermal management performance of phase

change material infiltrated microcellular metal foams, Appl Therm Eng. 64(2014) 147-154.

[36] W.S. Thomson, On the division of space with minimum partitional area, Acta Mathematica. 11(1971) 121-134.

[37] D. Weaire, R. Phelan, A counter-example to Kelvin's conjecture on minimal surfaces, Philosophical Magazine

Letters. 69(1994) 107-110.

[38] P. Kumar, F. Topin, Simultaneous determination of intrinsic solid phase conductivity and effective thermal

conductivity of Kelvin like foams, Applied Thermal Engineering. 71(2014) 536-547.

[39] L.J. Gibson, M.F. Ashby, Cellular Solids: Structures and Properties, Cambridge: Cambridge University Press.

1997

[40] E. Sadeghi, S. Hsieh, M. Bahrami, Thermal conductivity and contact resistance of metal foams, Journal of

Physics D Applied Physics. 44(2011) 125406.

[41] E. Takegoshi, Y. Hirasawa, J. Matsuo, K.I. Okui, A Study on Effective Thermal Conductivity of Porous Metals,

Nihon Kikai Gakkai Ronbunshu B Hen/transactions of the Japan Society of Mechanical Engineers Part B.

58(1992) 879-884.

[42] M. Fetoui, F. Albouchi, F. Rigollet, S.B. Nasrallah, Highly porous metal foams: Effective thermal conductivity

measurement using a photothermal technique, Journal of Porous Media. 12(2009) 939-954.

[43] J.W. Paek, B.H. Kang, S.Y. Kim, J.M. Hyun, Effective Thermal Conductivity and Permeability of Aluminum

Foam Materials, International Journal of Thermophysics. 21(2000) 453-464.

[44] E. Bianchi, T. Heidig, C.G. Visconti, G. Groppi, H. Freund, E. Tronconi, An appraisal of the heat transfer

properties of metallic open-cell foams for strongly exo-/endo-thermic catalytic processes in tubular reactors,

Chemical Engineering Journal. 198-194(2012) 512-528.


[45] R. Dyga, S. Witczak. Investigation of Effective Thermal Conductivity Aluminum Foams, Procedia

Engineering. 42(2012) 1088-1099.

[46] E.N. Schmierer, J. Paquette, A. Razani, K.J. Kim, Effective Thermal Conductivity of Fully-Saturated High

Porosity Metal Foam, ASME 2004 Heat Transfer/Fluids Engineering Summer Conference. 2004 229-237.

[47] W.Q. Xu, X.G. Yuan, L. Zhen, Study on effective thermal conductivity of metal foam matrix composite phase

change materials, Journal of Functional Materials. 40(2009) 1329-1332+1337.

[48] A. Bhattacharya, Thermophysical properties and convective transport in metal foam and finned metal foam

heat sinks, Ph.D. thesis. University of Colorado, Boulder, CO (2001).

[49] S. Feng, M. Shi, Y. Li, T.J. Lu, Pore-scale and volume-averaged numerical simulations of melting phase

change heat transfer in finned metal foam, International Journal of Heat & Mass Transfer. 90(2015) 838-847.

[50] Z. Chen, D. Gao, J. Shi, Experimental and numerical study on melting of phase change materials in metal

foams at pore scale, International Journal of Heat & Mass Transfer. 72(2014) 646-655.

[51] Z. Deng, X. Liu, C. Zhang, Y. Huang, Y. Cheng, Melting behaviors of PCM in porous metal foam

characterized by fractal geometry, International Journal of Heat & Mass Transfer. 113(2017) 1031–1042.

[52] X.H. Yang, X.Z. Meng, Z.N. Wang, L.W. Jin, Q.L. Zhang, Q.C. Zhang, T.J. Lu, Direct Numerical Simulation

on Melting Phase Change Behavior in Open-cell Metal Foam, Energy Procedia. 105(2017) 4254-4259.

[53] P. Zhang, Z.N. Meng, H. Zhu, Y.L. Wang, S.P. Pemg, Melting heat transfer characteristics of a composite

phase change material fabricated by paraffin and metal foam, Applied Energy. 2015.
[54] P. Zhang, Z.N. Meng, H. Zhu, Y.L. Wang, S.P. Peng, Experimental and Numerical Study of Heat Transfer
Characteristics of a Paraffin/Metal Foam Composite PCM, Energy Procedia. 75(2015) 3091-3097.
Figure captions

Fig. 1. Crystal cell structure based on the minimal surface energy theory.

Fig. 2. The formed skeleton structure of a metal foam and the extracted computational domain.

Fig. 3. Independent verification of mesh size.

Fig. 4. Effect of foam metal pore size on effective thermal conductivity of the composite.

Fig. 5. Comparison of simulated effective thermal conductivities and literature results.

Fig. 6. The effective thermal conductivity of composites comprised of different materials.

Fig. 7. Changes in the effective thermal conductivity with porosity of different foam metal composites.

Fig. 8. Phase and temperature distribution in the domain when liquid fraction = 0.25.

Fig. 9. Change in phase and temperature distribution with time at the location Z/L = 0.5.

Fig. 10. Temperature variations of monitoring points within the domain.

Fig. 11. Change in heat flux at the bottom surface of the domain.

Fig. 12. Change in liquid fraction over time.

Table captions

Table 1 Properties of materials used in the simulation.


(a) Crystal cell unit (b) skeleton structures

Fig.1.

(a) Skeleton structure (b) Computational domain

Fig.2.
14
12
10
λ W/(m·K)

8
6
4 foam metal/air
2 foam metal/water

mesh 1 mesh 2 mesh 3 mesh 4


Meshes

(a) The effective thermal conductivity simulation. (b) Melting process simulation

Fig.3.

14.4
Water
14.1 Air
W/(m·K)

13.8
13.5
13.2
12.9
12.6
10 20 30 40
PPI

Fig.4.
20 20
30

20 Xu [18]
10 10

 W/(m·K)
 W/(m·K)
 W/(m·K)

Bianchi [44] Zhang [29]


Fetoui [42] Simulation results Xiao [30]
Paek [43] Boomsma and Poulikakos[19] 10 Bhattacharya et al[22]
Takegoshi [41] Dul'nev[20] Xu [47]
0 Sadeghi [40] 0 Singh and Kasana[21]
E1[19]
Simulation results Schmierer [46]
Boomsma and Poulikakos[19]
0 Dul'nev[20]
Schmierer [46]
Dil'nev[20] Dyga and Witczak [45] Singh and Kasana[21]
Singh and Kasana[21] Bhattacharya [48] Simulation results
-10 -10 -10 B
0.80 0.85 0.90 0.95 1.00 0.80 0.85 0.90 0.95
 0.80 0.85 0.90 0.95 1.00
 

(a) Foam aluminum/Air (b) Foam aluminum/Water (c) Foam copper/Paraffin

Fig.5.

e=0.9097
s

0.1

0.01 0.1 1
f/s

Fig. 6.
35
N ic k e l fo a m /P a ra ffin
30 A lu m in u m fo a m /P a ra ffin
C o p p e r fo a m /P a ra ffin
25

 W/(m K)
20
15
10
5
0
0.84 0.88 0.92 0.96

Fig.7.

Z/L = 0.5 Z/L = 0.35 Z/L = 0.2 Z/L = 0.1 Z/L = 0.5 Z/L = 0.35 Z/L = 0.2 Z/L = 0.1
(a) Phase distribution. (b) Temperature distribution

Fig.8.
t=1s t=5s t = 10 s t = 15 s t=1s t=5s t = 10 s t = 15 s
(a) Phase distribution. (b) temperature distribution.

Fig.9.

350
point1
340 point2
point3
330
T/K

320

310

300
0 5 10 15 20 25
t/s

Fig.10.
16 16
330K
8 350K 8 Al
4 370K
Heat Flux (W)
Cu

Heat Flux(W)
2 4
Ni
1 2
0.5
1
0.25
0.125 0.5

0.0625 0.25
0 5 10 15 20 25 30 35 40
0 10 20 30 40 50 60
t/s
t/s

(a) Different heat source temperature. (b) Different foam metal composites.

Fig.11.

1.0 1.0
Al
1.0
Cu 330K
0.8 Ni 0.8 350K 0.8 =90.97%
=88.78%
Liquid Fraction

370K
Liquid Fraction

Liquid Fraction

0.6 0.6 0.6 =86.42%

0.4 0.4 0.4

0.2 0.2 0.2

0.0
0.0 0.0 0 5 10 15 20 25
0 5 10 15 20 25 30 35 40 0 10 20 30 40 50 60
t/s
t/s t/s

(a) Effect of metal matrix material (b) Effect of heating temperature (c) Effect of porosity of foam metal

Fig.12.
Table 1
Properties of materials used in this simulation.
Materials Paraffin Water Air Aluminum6101 Copper
Capacity kJ/(kg·K) 2340 4182 1006.43 963 377
Conductivity W/(m·K) 0.274 0.613 0.0265 218 398
Density kg/m3 770 998.2 1.225 2700 8440
Latent heat kJ/kg 260 - - - -
Thermal expansivity K -1 0.0011 - - - -
Melting point K 317 - - - -

Nomenclature
REV Representative Elementary Volume PCM Phase Change Material
TES Thermal Energy Storage BCC Body-Centered-Cubic
FCC Face-Centered-Cubic

Symbols
Г cross-sectional area q heat flow
L length T temperature
Ω heat flow direction ρ density
u velocity t time
P pressure μ viscocity
β thermal expansion coefficient g acceleration of gravity
c capacity λ thermal conductivity
A area Z height

Subscripts
liquidus solidification solidus melting
f filling material s solid material

You might also like