You are on page 1of 10

Solidification within a cavity filled with phase change

material

Lucas S. de Oliveira1✉ [0000-0001-7478-1477] and Kamal A. R. Ismail1[0000-0001-7671-1384]


1Energy Department, State University of Campinas, Faculty of Mechanical Engineering,
Mendeleiev street, 200, Cidade Universitária “Zeferino Vaz”, 13083-860, Barão Geraldo,
Campinas, Brasil

l235220@dac.unicamp.br,kamal@fem.unicamp.br

Abstract. Due to the serious impacts of global warming and pollution, coupled
with the imminent depletion of fossil fuels, the world is seeking to replace con-
ventional sources of energy with renewable alternatives. Among these alterna-
tives, solar energy and wind energy, although intermittent, are promising options.
To better harness the potential of these renewable sources, the development of
energy storage systems, particularly latent heat storage, is crucial to ensure reli-
able operation and increase the energy supply in renewable energy systems. This
article presents a numerical study of the solidification of phase change material
(PCM) within a rectangular cavity, considering variations in the cavity width.
The analyzed widths range from 10, 50, 100, 150, to 200 mm. The study uses a
one-dimensional model formulated based on the principles of heat conduction
and the enthalpy approach. The resulting system of equations is numerically
treated using the finite volume approximation. To assess the accuracy of the com-
putational model developed and implemented in Matlab, tests and comparative
validation were conducted against the established results in the field of study.
The proposed correlations for the solid-liquid interface position, which defines
the boundary between the solid and liquid phases of the PCM during the solidi-
fication process, show a remarkable agreement with the results of the reference
articles, with maximum deviations of 5.20% and 2.11%, respectively. The effects
of the vertical plate temperature and cavity width on the interface position, inter-
face velocity, stored energy, and time for complete solidification are presented
and discussed.

Keywords: Phase Change Material, Cavity, Interface Position.

1 Introduction

Given the increasing importance of implementing renewable energy sources to promote


environmental sustainability and mitigate greenhouse gas emissions, it is crucial to ad-
dress the challenges inherent in the effective use of energy from these sources. Among
such challenges, the intermittent nature of these energy sources stands out [1]. There-
fore, the continuous advancement of research and development becomes essential,
2

especially in thermal energy systems that rely on the principle of latent heat. In this
context, Phase Change Materials (PCM) emerge as a promising solution. These sub-
stances have the ability to store and release a significant amount of thermal energy
during transitions between physical states, such as fusion or solidification [2].
PCMs can be used to regulate the temperature of various systems and processes in-
volving heat exchanges. Their applications are diverse and include enhancing energy
efficiency in buildings, optimizing the performance of solar panels, temperature control
in food, providing thermal comfort in textiles, and extending the lifespan of electric
vehicle batteries [3].
The solidification of PCM is a complex process that is influenced by various factors,
such as boundary conditions, geometry, cavity dimensions, and the thermophysical
properties of the PCM, which include its thermal conductivity, specific heat capacity,
density, latent heat, and solidification temperature [4]. A highly relevant factor in the
design and optimization of systems using PCMs is the geometry of the container that
houses this material. Geometry has a significant impact on the thermal performance,
efficiency, and reliability of the system [5]. Currently, various geometries for PCMs
are being investigated, including flat, cylindrical, and spherical shapes, through various
analytical, approximate, and numerical methods [6].
A useful tool for analyzing the physical phenomena involved in PCM solidification
is numerical modeling, which can be performed using different numerical schemes such
as finite difference, finite element, or finite volume methods [7]. One of the most com-
monly used methods for modeling phase change problems is the enthalpy method. This
approach involves using an enthalpy function to represent the total heat content per unit
mass of the material under consideration, rather than limiting the analysis to tempera-
tures alone [8]. This approach simplifies the problem by eliminating the need to track
the position of the boundary between the solid and liquid phases of the material. As a
result, conventional numerical methods can be applied more easily and accurately.
The objective of this work is to develop a simple model to analyze the solidification
process of phase change materials (PCMs) within a cavity. The model uses the one-
dimensional energy equation and is numerically solved using enthalpy and finite vol-
ume methods. The numerical results are compared with literature data, demonstrating
good agreement.

2 Methodology

2.1 Problem formulation

The considered physical configuration involves the solidification process of a Phase


Change Material (PCM) within a cavity. Initially, the PCM is in a liquid state at an
initial temperature (T1) of 2°C. The left wall of the cavity is maintained at a constant
temperature (Tw) below the phase change temperature (Tm), thus initiating the solidi-
fication process, with the solidification front advancing in the positive direction of the
x-axis. The right side wall is considered thermally insulated, and the top and bottom
sides are infinitely long, with heat transfer occurring only in the x-direction, ensuring a
zero heat flux through these walls. In this thermal model, water is adopted as the PCM,
3

starting in the liquid state with characteristic physical parameters, such as a density of
994 kg/m³, thermal conductivity of 0.652 W/(m·K), and specific heat of 4174 J/kg·K
[9].
Fig. 1 shows the problem at hand, in which heat conduction predominates to simplify
the calculation model, while the effects of convection in the liquid PCM are neglected.
This can be described by the energy equation applied to the solid phase of the PCM.

Ts   Ts 
ρcs =  ks 
t x  x  (1)

For the liquid phase, it is possible to calculate using the following energy equation:

T T 
ρc l =   k l (2)
l t x  l x 

Fig. 1.Schematic diagram of the physical problem and coordinate system.

By the boundary conditions at the liquid/solid interface, as stated by Ozisik [10]:

Ts T
l = ρL  S  , T = T = T (3)
ks -k   s m
x l x  t  l

Where S is the instantaneous position of the interface between the solid and liquid
phases, and L is the latent heat. To define the problem, it is necessary to establish the
domain region, boundaries, and initial conditions. The temperature on the left side wall
is kept constant, serving as a boundary condition.

T = Tw (4)

The rate of heat loss from the bottom was zero due to perfect insulation. On the right
side wall, the boundary condition can be expressed mathematically as follows:
4

T
=0 (5)
x x=L

The initial and final conditions for the phase change material in the liquid state:

T ( x, t = 0)  Tm + ΔT (6)

The initial and final conditions for the phase change material in the solid state:

(
T x, t = t
f )  Tm - ΔT (7)

Where Δt represents half of the temperature range of the phase change, and tf indicates
the final moment of the phase change process. Following the work of Bonacina et al.
[8], the enthalpy per unit volume of the PCM is described as:

H ( T ) =  C ( T )dT + λδ ( T - Tm ) (8)
T

The specific heat capacity per unit volume of a material, denoted by C(T), the latent
heat per unit volume, denoted by λ, and the Dirac delta function, denoted by δ(T-Tm),
can be used to calculate the equivalent material specific heat capacity per unit volume.

dH ( T )
C (T) = = C ( T ) + λδ ( T - Tm ) (9)
dT

The expression representing the behavior of the specific heat equivalent of the PCM in
the solid phase is:

C ( T ) = Cs when T < Tm - ΔT (10)

For the liquid phase:

C ( T ) = C when T > Tm + ΔT (11)


l

The thermal conductivity of the PCM can be written as:

ks ( T ) T  Tm - ΔT

k (T) =  (12)
 l( )
 k T T  Tm + ΔT

By substituting Eq. (9) into Eq. (8), the specific heat capacity can be obtained by inte-
grating over the phase change temperature range as:
Tm +ΔT Tm Tm +ΔT
H (T) =  C ( T ) dT = λ +  Cs ( T ) dT +  C ( T ) dT (13)
Tm -ΔT Tm -ΔT Tm l

To facilitate numerical calculations, the variables were made dimensionless below:


5

T-T1 x ks t ΔT C k
θ= ; X= ; Fo= 2
;ε = ;C = ;k = (14)
Tm -T1 L Cs L Tm -Tw Cs ks

Following the same treatment as in [14] and putting the heat balance equations and the
associated boundary conditions in dimensionless form, one can write the final equations
in the form.
θ   θ 
C (θ ) = k (θ ) (15)

Fo X  X 
The boundary conditions are:

T
θ=0 , =0 (16)
X X=L

The dimensionless thermal capacity and the thermal conductivity were developed in a
similar manner as in [14] and are omitted for brevity. To solve this problem, a numerical
approach involving the discretization of differential equations was used. This entails
dividing the domain into control volumes. A portion of a one-dimensional grid is shown
in Fig. 2.

Fig. 2. Representation of the finite volume method.

The equation (15) was integrated in time and space. Therefore, similarly to [11], one
can obtain the energy equation for the control volumes within the domain in the fol-
lowing form.

a pθ p =a e fθe + (1-f ) θ e  +a w fθ w + (1-f ) θ w  - a e -a w × (1-f ) θ p +a po θp (17)


1 1 0 1 0 0 0

According to Patankar [11], it is possible to have an explicit formulation where f = 0,


an implicit formulation where f = 1, or a Crank-Nicolson formulation where f = 0.5.
The mathematical model and corresponding initial and boundary conditions were used
to develop a numerical code implemented in Matlab. The code underwent testing and
optimization. Various simulations were conducted with time intervals ranging from 10 -
1
to 10-6 s, with a value of Δt = 10-4 s was selected for good accuracy. The solution was
validated with longer time intervals, as illustrated in Fig. 3, and did not exhibit signifi-
cant differences in results compared to the smaller intervals, only an increase in com-
putation time. The domain region was discretized with different grid point quantities,
such as 20, 50, 100, 150, and 200, as shown in Figure 3. The value of 200 was chosen
for providing greater precision, with a Δx = 0.5.
6

Fig. 3. Optimization of results through time step adjustment.

Fig. 4. Optimization of results through grid points adjustment.

2.2 Model validation

The thermal model and solution method were validated against numerical data obtained
by Rubinsky [11], as shown in Fig. 5. The author described a numerical study con-
ducted to investigate the solidification of a saline solution subjected to a temperature of
-20 °C. The study used the finite element method to track the interface, considering the
Mullins-Sekerka stability criteria. The validation shows a maximum difference of
5.20%.
The numerical results obtained by Voller [12] were compared to those of this work
in Fig. 6, considering the PCM as water at an operating temperature of -10 °C. The
7

good agreement between the current numerical predictions and those of Voller [12] can
be observed, with a maximum difference of 2.11%.

Fig. 5. Validation of the interface position with [14].

Fig. 6. Validation of the interface position with [15].

3 Results and Discussion

Fig. 7 shows the effect of the plate low temperature on the phase change interface po-
sition. This behavior is due to the increase of the thermal gradient between the wall and
the liquid PCM resulting in greater heat transfer and increase of the interface position.
The relationship between wall temperature and interface velocity is demonstrated in
Fig. 8. As the wall temperature decreases, the temperature gradient between the cold
wall and the PCM increases. This results in an increase in heat transfer rate, solidifica-
tion rate, and interface velocity.
8

Fig. 7. Effect of plate temperature on the interface position.

Fig. 8. Effect of plate temperature on the interface velocity.

Fig. 9 shows that the stored energy increases as the plate temperature decreases, which
increases the temperature difference between the cold plate and the liquid PCM. This
results in an increase in the solidified mass (stored latent heat) due to the phase change,
and the stored sensible heat due to temperature variation, contributing to the total en-
ergy of the system.

Fig. 9. Effect of plate temperature on the stored energy.


9

Fig. 10 shows the total amount of sensible heat stored, resulting from the sum of sensi-
ble heat in the liquid phase and sensible heat in the solid phase. As the plate's tempera-
ture decreases, the stored sensible energy increases, leading to a greater release of heat
due to an increase in the temperature difference between the plate and the PCM.

Fig. 10. Variation of sensible heat with temperature.

Fig. 11 shows the influence of cavity width on the total solidification time. It is evident
that the time required for complete solidification increases as the width of the cavity
expands. This increase is directly related to the increase of thermal resistance between
the cold wall and the PCM. Additionally, it can be observed that the time to achieve
complete solidification decreases as the wall temperature is reduced, as previously ex-
plained.

Fig. 11. Effect of cavity width on the complete solidification time.


10

4 Conclusion

As found the reduction of temperature on the cold wall results in an increase in phase
change parameters, such as the position and velocity of the interface, as well as the
stored energy. Additionally, the width of the cavity has direct influence on the increase
in the total solidification time.

Acknowledgment. The authors wish to thank the Foundation for Research and Technological
Development of Maranhão (FAPEMA) for the Master scholarship offered to the first author.

References
1. Suberu MY, Mustafa MW, Bashir N (2014) Energy storage systems for renewable energy
power sector integration and mitigation of intermittency. Renewable and Sustainable Energy
Reviews 35: 499-514.doi:10.1016/j.rser.2014.04.009.
2. Zhang H, Baeyens J, Caceres G, Degreve J, Lv Y (2016) Thermal energy storage: recent
developments and practical aspects. Prog. Energy Combust. Sci. 53:1–40.doi:
10.1016/j.pecs.2015.10.003.
3. Ismail KA, Lino FA, Machado PLO, Teggar M, Arıcı M, Alves TA, Teles MP (2022) New
potential applications of phase change materials: A review. Journal of Energy Storage 53:
105202.doi:10.1016/j.est.2022.105202.
4. Kenisarin MM (2014) Thermophysical properties of some organic phase change materials
for latent heat storage. A review. Solar Energy 107: 553-575.doi:
10.1016/j.solener.2014.05.001.
5. Choure BK, Alam T, Kumar R (2023) A review on heat transfer enhancement techniques
for PCM based thermal energy storage system. Journal of Energy Storage 72: 108161.doi:
10.1016/j.est.2023.108161.
6. Liu S, Li Y, Zhang Y (2014) Mathematical solutions and numerical models employed for
the investigations of PCMs phase transformations. Renewable and Sustainable energy re-
views 33: 659-674.doi: 10.1016/j.rser.2014.02.032.
7. Dolado P, Lazaro A, Marin JM, Zalba B (2011) Characterization of melting and solidifica-
tion in a real scale PCM-air heat exchanger: Num. model and experimental validation. En-
ergy Conversion and Management 52(4):1890-1907.doi:10.1016/j.enconman.2010.11.017
8. Bonacina C, Comini G, Fasano A, Primicerio M (1973) Numerical solution of phase-change
problems. Int J Heat Mass Transfer 16(10):1825-1832.doi:10.1016/0017-9310(73)90202-0
9. Holman, J.P (2010) Heat transfer. Macgraw-Hill, New York.
10. Hahn DW, Özisik MN (1993) Heat conduction. John Wiley & Sons, New York.
11. Patankar SV (1982) N. Heat Transfer and Fluid Flow. Hemisphere Publishing, New York.
12. Rubinsky B (1983) Solidification processes in saline solutions. J Cryst Growth 62(3):513-
522. doi:10.1016/0022-0248(83)90394-9.
13. Voller V, Cross M (1981) Accurate solutions of moving boundary problems using the en-
thalpy method. Int J Heat Mass Transfer 24(3):545-556.doi:10.1016/0017-9310(81)90062-
4.
14. Ismail KA, Gonçalves MM, Lino FA (2016) Solidification of PCM around a finned tube:
Modeling and experimental validation. J Basic Appl Res Int 12(2):115-128.

You might also like