You are on page 1of 30

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/308092864

CFD simulation of melting process of phase change materials (PCMs) in a


spherical capsule

Article in International Journal of Refrigeration · September 2016


DOI: 10.1016/j.ijrefrig.2016.09.007

CITATIONS READS

50 1,338

4 authors:

H. Sattari Ali Mohebbi


Shahid Bahonar University of Kerman Shahid Bahonar University of Kerman
2 PUBLICATIONS 70 CITATIONS 148 PUBLICATIONS 1,872 CITATIONS

SEE PROFILE SEE PROFILE

M. Mehdi Afsahi Amir Azimi Yancheshme


Shahid Bahonar University of Kerman Sharif University of Technology
27 PUBLICATIONS 214 CITATIONS 13 PUBLICATIONS 172 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Multi gene genetic programming View project

Molecular Dynamics Simulation View project

All content following this page was uploaded by Ali Mohebbi on 19 October 2018.

The user has requested enhancement of the downloaded file.


CFD simulation of melting process of phase change materials (PCMs) in a

spherical capsule

H. Sattaria, A. Mohebbia,*, M. M. Afsahia, A. Azimi Yancheshmeb


a
Department of Chemical Engineering, Faculty of Engineering, Shahid Bahonar
University of Kerman, Kerman, Iran
b
Chemical and Petroleum Engineering Department, Sharif University of
Technology, Tehran, Iran
*Corresponding author: Tel & Fax: +983432118298

E-mail: amohebbi2002@yahoo.com, amohebbi@uk.ac.ir

1
Abstract

The present study is focused on CFD simulation of constrained melting of

Phase Change Materials (PCMs) in a spherical container. To investigate the

melting process of the PCM, its melting fraction was analyzed at different

times. The results indicated the existence of thermally stable layers on the top of

the sphere. Moreover, inspection of the calculated temperatures at different

points along the vertical axis indicates the existence of some disturbances at the

bottom of the sphere due to the natural convection. After the validation of the

results, the effects of different parameters such as the surface temperature of the

capsule, the initial temperature and the size of the spherical capsules, on the

melting process were investigated. The initial temperature did not affect the

melting rate, whereas melting rate increased by increasing the surface

temperature of the capsule and also decreasing the diameter of the sphere. The

results showed that the surface temperature of capsule compared to geometrical

parameters and other operational conditions can have a greater influence on the

melting rate and the heat flux.

Keywords: Constrained melting; Phase change materials; Computational fluid

dynamics (CFD); Spherical container

2
1. Introduction

Recently, latent heat thermal energy storage systems have been widely

applied in cool storage for central air-conditioning, hot storage for air heating,

solar energy, energy efficient buildings and waste heat recovery. Storage

systems allow for the efficient and rational utilization of the available resources

or renewable energies, by using the time lag between production, as well as the

availability of the energy and its consumption in the demanding systems. Phase

change materials (PCM) are mainly used to provide higher storage densities. As

a serious problem PCMs have relatively low thermal conductivities. Therefore,

increasing the surface/volume ratio of PCMs in order to increase the heat-

transfer rate is very appealing. This can be done by packing a volume with a

great number of PCM capsules. The spherical geometry is one interesting case

for heat storage applications, since spheres are much employed in packed beds.

Due to the complexity of such systems, it is often more effective to first model

the behavior of an individual sphere and then describe it with a simple

parametric model in the packed bed modeling.

Roy and Sengupta [1] examined the melting process with the solid phase

initially uniformly super cooled. In order to include the effects of a temperature

gradient in the solid core, they modified the heat transfer equation. At each time

step, the unsteady conduction equation has been solved numerically using a

toroidal coordinate system with a suitable immobilization of the moving

3
boundary to transform the infinite domain into a finite one. In another paper [2],

they investigated the outcome of natural convection on the melting process. In

order to reduce the computational effort and the time, they made some

simplifications. They obtained the non-dimensional melting time and the heat

transfer coefficient as functions of the PCM property values, the operating

temperature and the physical size.

Ettouney et al. [3] experimentally evaluated the heat transfer (during

energy storage) and release for the phase change of paraffin wax in the spherical

shells. They showed that an increment in the Nusselt number of the sphere with

a larger diameter is attributed to the increase in the natural convection cells in

the PCM inside the sphere. The natural convection role is enhanced upon an

increase in the sphere diameter and the air temperature. On the other hand,

during solidification the wax layers are formed inward, nucleating on the sphere

walls. As solidification progresses, the melt volume becomes smaller and the

role of natural convection diminishes rapidly.

Khodadadi and Zhang [4] also considered the effects of buoyancy-driven

convection on the constrained melting of phase change materials within

spherical containers. They reported that during the initial melting process, the

conduction mode of the heat transfer is dominant. As the buoyancy-driven

convection is strengthened due to the increase in the melting zone, the process

at top region of the sphere is much faster than at the bottom due to the

increment of the conduction mode of heat transfer. They found that buoyancy-

4
driven convection speeds up the melting process compared to the diffusion-

controlled melting.

Assis et al. [5] numerically and experimentally explored the melting

process of a PCM in a spherical geometry. The results of their experimental

investigation, including visualization, were favorably comparable with the

numerical results and therefore suitable for validation of the mathematical

approach. Their computational results showed how the transient phase-change

process depends on the thermal and geometrical parameters of the arrangement.

Veerappan et al. [6] investigated the phase change behavior of 65 mol% capric

acid and 35 mol% lauric acid, calcium chloride hexa hydrate, n-octadecane, n-

hexadecane, and n-eicosane inside the spherical containers to identify a suitable

heat storage material. They created analytical models for solidification and

melting of PCM in a spherical shell with conduction, natural convection, and

heat generation and found a good agreement between the analytical predictions

and the experimental data. Both models were validated with the experimental

work of Eames and Adref [7].

Regin et al. [8] examined the heat transfer performance of a spherical

capsule using paraffin wax as PCM placed in a convective environment during

the melting process. The model results were in a good agreement with the

experimental data.

Tan [9] investigated the melting of the phase change material (PCM)

inside a sphere using n-octadecane for both constrained and unconstrained

5
melting processes. For constrained melting, paraffin wax (n-octadecane) was

immobilized, through the use of thermocouples when melting is done inside a

transparent glass sphere. Their experimental setup provided a detailed

temperature data that were gathered along the vertical diameter of the sphere

during the melting process. Tan and Khodadadi [10] also experimentally and

numerically investigated the constrained melting of PCMs inside a spherical

capsule to understand the role of the buoyancy-driven convection.

In this study, a computational fluid dynamics (CFD) modeling on the

constrained melting of phase change material (PCM) in a spherical container

was performed. The effects of different parameters like the capsule size, the

surface temperature, the initial temperature and the Stefan number, on the PCM

melting process were investigated.

2. CFD simulation

CFD simulation was carried out using the commercial Fluent 6.3

software. The physical model and computational procedure are discussed below.

2.1. Physical model

Consider a spherical shell with an internal radius (Ri) of 50.83 mm and a

wall thickness of 1.5 mm initially filled with a solid PCM at an initial

temperature (Ti) lower than the melting temperature (Tm).The schematic

6
diagram of the physical model is shown in Fig. 1. At t (time) > 0, a constant

temperature (T0) greater than the melting temperature is imposed on the surface

of the sphere, i.e. T0>Tm. Melting starts at the surface, moving the solid-liquid

interface towards the middle of the field. Since the present work is focused on

the analysis of constrained melting, both the solid and liquid phases have the

same density.

The container was made of glass with a thermal conductivity of 1.14

Wm-1K-1. The properties of the PCM, based on n-octadecane as a commercially

available material, are given in Table 1.

Fig.1.The computational domain

7
Table 1. Thermophysical properties of n-octadecane.

Melting temperature 28.2 oC

Density 772 kgm-3

Kinematic viscosity 5 x 10-6 m2s-1

Specific heat 2330 Jkg-1K-1

Thermal conductivity 0.1505 Wm-1K-1

Latent heat of fusion 243.5 kJkg-1

Thermal expansion 0.00091 K-1

coefficient

For the experimental data that we used [9], the initial temperature was

1°C below the melting temperature and the surface temperature was 40°C. The

Prandtl, Stefan and Rayleigh numbers for this problem were 59.6, 0.113, and

2.63×108, respectively.

2.2. Computational procedure

The numerical approach makes it possible to predict specifications of the

melting process that take place inside the sphere. The flow was considered

unsteady, laminar, incompressible and two-dimensional. In order to simulate the

constrained melting, it is supposed that both the liquid and the solid phases are

isotropic, homogeneous and remain in thermal equilibrium at the interface. For

8
the phase-change region inside the PCM, enthalpy-porosity approach [11-12]

was used. In this technique, the melt interface is not tracked explicitly.

Alternatively, a quantity called the liquid fraction, indicating the fraction of the

cell volume that is in liquid form, is associated with each cell in the domain. In

this way, the governing equations for the PCM are as follows:

Continuity:

𝜕𝜌 (1)
+ ∇. (𝜌𝑉) = 0
𝜕𝑡

Momentum:

𝜕(𝜌𝑉) (2)
+ ∇. (𝜌𝑉) = −∇𝑃 + 𝜇∇2 𝑉 + 𝜌𝑔 + 𝑆
𝜕𝑡

Energy:

𝜕 (3)
(𝜌𝐻) + ∇. (𝜌𝑉𝐻) = ∇. (𝑘∇𝑇)
𝜕𝑡

where V is the fluid velocity vector, ρ is the density, μ is the dynamic viscosity,

P is the pressure, g is the gravitational acceleration, k is the thermal conductivity

and H is the specific enthalpy, which is defined as the sum of the sensible

enthalpy (h) and the latent heat (ΔH):

𝐻 = ℎ + ∆𝐻 (4)

where

9
𝑇 (5)
ℎ = ℎ𝑟𝑒𝑓 + ∫ 𝐶𝑝 𝑑𝑇
𝑇𝑟𝑒𝑓

The content of the latent heat can vary between zero (for a solid) and L (for a

liquid):

∆𝐻 = 𝛽𝐿 (6)

The liquid fraction (β) can be written as:

𝛽=0 𝑖𝑓 𝑇 < 𝑇𝑠𝑜𝑙𝑖𝑑𝑢𝑠


𝛽=1 𝑖𝑓 𝑇 > 𝑇𝑙𝑖𝑞𝑢𝑖𝑑𝑢𝑠
𝑇 − 𝑇𝑠𝑜𝑙𝑖𝑑𝑢𝑠 (7)
𝛽= 𝑖𝑓 𝑇𝑠𝑜𝑙𝑖𝑑𝑢𝑠 < 𝑇 < 𝑇𝑙𝑖𝑞𝑢𝑖𝑑𝑢𝑠
{ 𝑇𝑙𝑖𝑞𝑢𝑖𝑑𝑢𝑠 − 𝑇𝑠𝑜𝑙𝑖𝑑𝑢𝑠

By definition, solidus and liquidus temperatures are the maximum temperature

at which the material is at solid state and minimum temperature at which the

material is at liquid state, respectively. However, for melting of a pure

substance, phase change occurs at a distinct melting temperature. Therefore,

solidus and liquidus temperatures are the same. In order to determine the liquid

fraction at mushy zone, according to the method was given by Voller and

Prakash[11], a tiny temperature interval Δ𝑇 near the melting point temperature

(Tm) was considered and assumed that liquefaction occurs at a certain range of

temperature (𝑇𝑚 − Δ𝑇 and 𝑇𝑚 + Δ𝑇), as shown in Fig. 2. Consequently, using

Equation (7) and by substituting 𝑇𝑚 − Δ𝑇 and 𝑇𝑚 + Δ𝑇 for 𝑇𝑠𝑜𝑙𝑖𝑑𝑢𝑠 and

𝑇𝑙𝑖𝑞𝑢𝑖𝑑𝑢𝑠 respectively, liquid fraction was calculated.

10
Fig. 2. The relation between enthalpy and temperature of a pure substance.

The source term S in the momentum equation is the Darcy's law damping

term that is added to the momentum equation due to the phase change effect on

the convection. That is given by:

𝐶(1 − 𝛽)2 (8)


𝑆= 𝑉
𝛽3

The coefficient C is a mushy zone constant. This constant measures the

amplitude of the damping. The higher this value, the steeper is the transition of

the velocity of the material to zero as it solidifies. Values between 10 4 and 107

are the recommended values for most calculations [12] and values exceeding

this range may cause the solution to oscillate. In the present research, different

values of the constant C were investigated to examine the effect of the mushy

zone constant.

In order to handle pressure-velocity coupling, SIMPLE (semi-implicit

method for pressure-linked equations) algorithm of Patankar [13] was


11
employed. The PRESTO scheme was used for the pressure correction equation.

This scheme uses the discrete continuity balance for a “staggered” control

volume about the face to compute the “staggered” pressure [13]. For flows with

natural convection, rotating flows, and flows in strongly curved domain, using

PRESTO scheme is recommended [12]. Other equations (i.e. momentum and

energy) were discretized using power law scheme, which interpolates the value

of a variable using the exact solution to a one-dimensional convection-diffusion

equation. The under-relaxation factors for the velocity components, thermal

energy, pressure correction and liquid fraction were 0.1, 0.3, 1 and 0.9,

respectively.

Different grid densities were tried to make sure that the solution is independent

from the adopted grid density (see Fig. 3). As a result, 7200 grids were found

sufficient for the present study. Adopting a fine grid distribution in the radial

direction allows for the use of longer time steps. The time required for

achieving the full melting is a good measure of time step dependence. By

comparing the selected quantities obtained from simulations using 0.05, 0.1,

0. 5 and 1 s, the time step for integrating the time derivatives was set to 1 s.

12
1
0/9
0/8
0/7
Liquid fraction

0/6
0/5
0/4
0/3 12800 node

0/2 5000 node

0/1 7200 node

0
0 20 40 60 80 100 120 140
Time(min)

Fig. 3. Variations of liquid fraction vs. time at different grid sizes.

3. Results and discussion

The effect of different values of the mushy zone constant on the melt

fraction with a Stefan number of 0.132 and a shell diameter of 101.66 mm, is

shown in Fig. 4. As one can see from this figure, the value of 107 is suitable for

the present simulation where the numerical result is in most accordance with the

experimental data. Fig. 4 also confirms the validation of the result of CFD work.

13
Fig. 4. Variations in the liquid fraction vs. time at different mushy zone

constants (Ste= 0.113 and D=101.66 mm).

The colored contours of solid-liquid front of the PCM at various time

instants for constrained melting process at a wall temperature of 40 °C and an

initial temperature of 27.2 °C are shown in Fig.5. The blue and red colors

represent the solid and melted parts, respectively. In the beginning of the

process, the outer surface of the solid PCM is in contact with the inner wall of

the sphere so that heat conduction dominates between the solid PCM and the

wall. This makes the formation of a thin layer of liquid between the solid PCM

and the wall of the sphere. As time progresses, the molten zone expands and the

liquid layer grows. The liquid film is then forced up on the interior surface of

14
the sphere, occupying the region above the solid PCM. By moving towards the

top part of the sphere, the warm liquid is replaced with the cold liquid.

Therefore, natural convection in combination with a hot rising curved wall jet

are dominant at the top and side regions of the sphere.

After 40 and 60 minutes, as can be seen in Fig. 5, more intense melting in

the upper zone compared to the side and bottom sections is observed. This

phenomenon is attributed to the strong role of natural convection in the upper

zone of the sphere. After 40, 60 and 80 minutes, a waviness at the bottom part

of the PCM is observed. This can be attributed to the existence of stable and

unstable thermal layers at the bottom part of the sphere. This result as well as

the observed shape of the solid, are in good agreement with the experimental

data in the literature [9].

15
0 min 20 min 40 min

60 min 80 min 100 min

120 min

Fig. 5. Constrained melting phase front at wall temperature of 40 °C and

initial temperature of 27.2 °C at different times.

The temperature contours of the right side of the sphere as well as the

streamlines of the left side of the sphere, at a wall temperature of 40 °C for a

sphere of 101.66 mm diameter, are shown in Fig. 6. At 20 and 40 minutes,

formation of recirculating vortexes between the solid PCM and the inner wall of

the sphere as a result of the replacement of the warm liquid with cold liquid at

the bottom part of the sphere is obvious. It can also be seen that more melting
16
occurs at the upper part of the sphere due to the contribution of natural

convection (see Fig. 5). The whole PCM is melted after 120 minutes.

20 min 40 min

17
60 min 80 min

100 min 120 min

Fig. 6. Detailed temperature contours (T in Kelvin) and streamlines for a

wall temperature of 40 °C at different times.

Comparison of experimental temperatures with the results of CFD

simulation for three points A, B, and C along the vertical diameter of the sphere

(see Fig.1) is shown in Fig. 7.

18
42 40
40
38
38
36
Temperature (°c)

Temperature(°C)
36
34
34
32
32
30 30

28 28
present work present work
26 26
Experimental data[9] Experimental data[9]
24 24
0 20 40 60 80 100 120 140 0 20 40 60 80 100 120 140
Time (min) Time(min)

(a) (b)

40
38
36
Temperature(°C)

34
32
30
28
present work
26
Experimental data[9]
24
0 20 40 60 80 100 120 140
Time(min)

(c)

Fig. 7. Comparison of the computed and measured temperatures at (a)point A at

a distance of 47.8 mm below the center (b) point B at a distance of 37.5 mm

below the center (c) point C at a distance of 25 mm above the center.

The disordered temperature data at points A and B, show the chaotic

convection motion in the unstable liquid layer at the bottom of the sphere. The

computed temperatures at point C (located on the upper side of the sphere) are

19
in a good agreement with the experimental data. This is due to the stable nature

of the liquid layer at the upper part of the sphere.

To prove the existence of liquid layer where the point A is located, the

variation of liquid fraction at the point A during melting time was calculated

(see Fig. 8). The results revealed that the PCM was at a transition state from

solid to liquid during 7.5 to 9.5 minutes and after 9.5 minutes it was completely

at liquid state. Therefore, the high frequency oscillation in the temperature

versus time may be found at this point.

1
Liquid fraction at point A

0/8

0/6

0/4

0/2

0
7 7/5 8 8/5 9 9/5 10
Time (min)

Fig. 8. Changes in liquid fraction versus time at point A at distance of 47.8 mm


below the center.

The effect of initial temperature is showed in Fig. 9. It is observed that

different initial temperatures do not have a substantial effect on the melting

20
process. The melting rates for initial temperatures of 18.2 °C and 8.2 °C are the

same but the initial temperature of 27.2 °C has a few faster melting rate.

0/8

0/6
Liquid fraction

0/4
initial temperature at 8.2 °C

0/2 initial temperature at 27.2 °C


initial temperature at 18.2 °C
0
0 20 40 60 80 100 120 140
Time(min)

Fig. 9. Variations in the liquid fraction versus time at different initial

temperatures.

Fig. 10 shows variations in the liquid fraction with time at different

Stefan numbers. The Stefan number (Ste) is the ratio of sensible to latent heat of

the PCM. The surface temperature (TS) directly affects the value of the Stefan

number:

𝐶𝑝𝑙 (𝑇𝑆 − 𝑇𝑚 )
𝑆𝑡𝑒 = (9)
𝐿

here Cpl is the specific heat of the liquid PCM and L and Tm are the latent heat

and melting temperature of the PCM, respectively. The present simulation was

21
performed at three different Stefan numbers, namely Ste= 0.065 (Ts= 35 °C),

Ste= 0.113 (Ts= 40 °C) and Ste= 0.161 (Ts=45°C). By increasing the surface

temperature, the Stefan number will be increased whereas the time required for

completing the melting will be decreased.

0/8
Liquid fraction

0/6

0/4
ste=0.065(Ts=35°C)

0/2 ste=0.113(Ts=40°C)

ste=0.161(Ts=45°C)
0
0 30 60 90 120 150 180 210 240
Time(min)

Fig.10. Variations in the liquid fraction versus time at different surface

temperatures.

The effects of different shell diameters and Stefan numbers on the changes in
the PCM liquid fraction with time are shown in Fig.11.The slopes of these
graphs represent the melting rates. The highest melting rates belong to the
beginning of the melting process because of the heat conduction between the
inner wall of the shell and the solid PCM. As expected, for each shell diameter
the liquid fraction increases more rapidly when the Stefan number (or surface
temperature, Ts) is larger. Also, for each Stefan number, full melting is
accomplished more rapidly in a shell with a smaller diameter. However, the
effect of these parameters (i.e. diameter and surface temperature or Stefan
number) are not the same on the melting time and rate. With respect to the data
of Fig. 11, it can be extracted that, 28% increase in surface temperature (from
22
Ts = 35o C to Ts = 45o C) results in about 170% to 198% decrease in melting
time for each constant diameter. While 27% increase in diameter of container
(from D = 80 mm to D = 101.66 mm) results in about 80% increase in melting
time. Therefore, the effect of surface temperature on melting time as well as
melting rate is more than the effect of container’s diameter on these parameters.

23
1 1

0/8 0/8
Liquid fraction

Liquid fraction
0/6 0/6

0/4 0/4
Ste=0.065 Ste=0.065
0/2 Ste=0.113 0/2 Ste=0.113
Ste=0.161 Ste=0.161
0 0
0 20 40 60 80 100 120 140 0 20 40 60 80 100 120 140 160 180 200
Time(min) Time(min)

(a) D=60 mm (b) D=80 mm

0/8
Liquid fraction

0/6

0/4
Ste=0.065
0/2 Ste=0.113
Ste=0.161
0
0 30 60 90 120 150 180 210 240
Time(min)

(c) D=101.66 mm

Fig.11. Changes in the liquid fraction versus time at different Stefan numbers

and different shell diameters: (a) D=60 mm (b) D=80 mm (c) D=101. 66 mm.

Fig.12 indicates variations in the heat flux versus time at different shell

diameters and Stefan numbers. In all plots, a sharp drop in the heat flux is

observed at the beginning of the melting process but after a few minutes, it

decreases smoothly. When the whole PCM is melted, the heat flux almost

reaches zero. The results show that by increasing the Stefan number, the heat
24
7000 8000
6000 7000

5000 6000

Heat flux(Wm-2)
Heat flux(Wm-2)

4000 5000
4000
3000
3000
2000 ste=0.065 ste=0.065
ste=0.113 2000
ste=0.113
1000 ste=0.161 1000 ste=0.161
0
0
0 50 100 150
0 50 100 150 200
Time(min) Time(min)
(a) D=60 mm (b) D=80 mm
7000

6000

5000
Heat flux(Wm-2)

4000

3000
ste=0.065
2000
ste=0.113
1000 ste=0.161
0
0 50 100 150 200 250
Time(min)
(c) D=101.66 mm

Fig.12. Variations in the heat flux at different Stefan numbers and shell
diameters: (a) 60 mm (b) 80 mm (c) 101.66 mm.

transfer rate increases. Integration of the area under the heat flux graphs in

Fig.12 gives the total amount of heat collection. As expected, by increasing the

diameter of the sphere and due to the increase in the amount of the PCM, the

heat collection is enhanced.

25
4. Conclusions

The numerical study of the constrained melting process of n-octadecane

paraffin wax as a PCM in a spherical container at different operating conditions

and various diameters of the container was successfully done. Detailed phase

and flow fields were provided inside the system by simulations and the results

were discussed and compared with the experimental data. At the beginning, the

heat conduction to the PCM is dominant but after formation of a thin layer of

melted liquid between the inner wall and the solid PCM, natural convection

becomes dominant at the top half of the sphere. The comparison between the

calculated and measured temperatures at the top half of the sphere clearly shows

the stable nature of the molten liquid layer in that zone. The unstable liquid

layer at the bottom of the sphere is responsible for the waviness of the PCM at

this zone. Based on the simulation results, change in surface temperature has the

most deterministic effect on melting rate, melting time and the rate of heat

transfer in comparison with other parameters such as diameter of container and

initial temperature of PCM.

26
Nomenclature

Cpl Specific heat of liquid PCM, Jkg-1 K-1

H Specific enthalpy, Jkg-1

k Thermal conductivity, Wm-1 K-1

L Latent heat, Jkg-1

P Pressure, Pa

Ri Internal radius, m

t Time, s

Ti Initial temperature, K

Tm Melting temperature, K

Ts Surface temperature, K

V Velocity, ms-1

Greek symbols

 Liquid fraction, dimensionless

 Density, kgm-3

27
µ Dynamic viscosity, kgm-1s-1

References

[1] Roy SK, Sengupta S. Melting of a free solid in a spherical enclosure: effects

of subcooling. J Solar Energy Eng. 1989; 111:32–6.

[2] Roy SK, Sengupta S. Gravity-assisted melting in a spherical enclosure:

effects of natural convection. Int J Heat Mass Transfer 1990; 33:1135–47.

[3] Ettouney H, El-Dessouky H, Al-Ali A. Heat transfer during phase change of

paraffin wax stored in spherical shells. ASME J Solar Energy Eng. 2005; 127:

357–65.

[4] Khodadadi JM, Zhang Y. Effects of buoyancy-driven convection on melting

within spherical containers. Int J Heat Mass Transfer 2001; 44:1605–18.

[5] Assis E, Katsman L, Ziskind G, Letan R. Numerical and experimental study

of melting in a spherical shell. Int J Heat Mass Transfer 2007; 50 (9-10): 1790–

804.

[6] Veerappan M, Kalaiselvam S, Iniyan S, Goic R. Phase change characteristic

study of spherical PCMs in solar energy storage. Solar Energy 2009; 83: 1245–

52.

[7] Eames IW, Adref KT. Freezing and melting of water in spherical enclosures

of the type used in thermal (ice) storage systems. ApplTherm Eng. 2002; 22:

733–45.
28
[8] Regin AF, Solanki SC, Saini JS. Experimental and numerical analysis of

melting of PCM inside a spherical capsule. In: Proceedings of the 9th

AIAA/ASME joint thermos physics and heat transfer conference (CD ROM).

Paper AIAA 2006-3618; 2006. p. 12.

[9] Tan FL. Constrained and unconstrained melting inside a sphere. Int

Commun Heat Mass Transfer 2008; 35:466–75.

[10] Tan FL, Hosseinizadeh SF, Khodadadi JM, Fan L. Experimental and

computational study of constrained melting of phase change materials (PCM)

inside a spherical capsule. Int J Heat Mass Transfer 2009; 52:3464–72.

[11] Voller V, Prakash C. A fixed grid numerical modeling methodology for

convection-diffusion mushy region phase-change problems. Int J Heat Mass

Transfer 1987; 30:1709–19.

[12] Fluent 6.3.22 Users’ Guide, Fluent, Inc., 2006.

[13] Patankar S. V. Numerical Heat Transfer and Fluid Flow. Hemisphere,

Washington, DC. 1980.

29

View publication stats

You might also like