Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

Numerical Heat Transfer, Part B: Fundamentals

An International Journal of Computation and Methodology

ISSN: (Print) (Online) Journal homepage: www.tandfonline.com/journals/unhb20

Heat transfer analysis during the solidification of


RT82 paraffin in big-scale metal foam-based latent
thermal storage unit

Atef Chibani, Slimane Merouani, Mohamed Razi Morakchi, Noureddine


Gherraf, Aissa Dehane, Ghania Mecheri, Cherif Bougriou & Djemaa
Guerraiche

To cite this article: Atef Chibani, Slimane Merouani, Mohamed Razi Morakchi, Noureddine
Gherraf, Aissa Dehane, Ghania Mecheri, Cherif Bougriou & Djemaa Guerraiche (2023) Heat
transfer analysis during the solidification of RT82 paraffin in big-scale metal foam-based latent
thermal storage unit, Numerical Heat Transfer, Part B: Fundamentals, 84:6, 794-815, DOI:
10.1080/10407790.2023.2222907

To link to this article: https://doi.org/10.1080/10407790.2023.2222907

Published online: 21 Jun 2023.

Submit your article to this journal

Article views: 121

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=unhb20
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS
2023, VOL. 84, NO. 6, 794–815
https://doi.org/10.1080/10407790.2023.2222907

Heat transfer analysis during the solidification of RT82 paraffin


in big-scale metal foam-based latent thermal storage unit
Atef Chibania,b , Slimane Merouania, Mohamed Razi Morakchic, Noureddine Gherrafd,
Aissa Dehanea, Ghania Mecherib, Cherif Bougrioue, and Djemaa Guerraichef
a
Laboratory of Environmental Process Engineering, Department of Chemical Engineering, Faculty of Process
Engineering, University Salah Boubnider Constantine 3, Constantine, Algeria; bFaculty of Sciences and Applied
Sciences, Department of Process Engineering, Larbi Ben M’hidi University, Oum El Bouaghi, Algeria; cIndustrial
Maintenance (LGE) Electrical Engineering Laboratory, Electrical Engineering Department, Faculty of
Technology, University of M’sila, BP, M’sila, Algeria; dLaboratory of Natural Resources and Management of
Sensitive Environments, Faculty of SE/SNV, Larbi Ben M’Hidi University, Oum El Bouaghi, Algeria; eMechanical
Engineering Department, Faculty of Technology, University of Mostefa Ben Boulaid, Batna, Algeria 2, Batna;
f
Applied Energy Physics Laboratory (LPEA) Department of Physics Faculty of Matter Sciences, University of
Batna, Batna, Algeria

ABSTRACT ARTICLE HISTORY


Metal foam (MF) and nano-sized particles (NSP) are regarded as patent Received 3 November 2022
tools for enhancing the thermal performance of phase change materials Revised 21 April 2023
(PCM)-based latent thermal energy storage unit (LTES), but data on this Accepted 4 June 2023
issue for large-scale installations is very scarce. This study provides a com-
KEYWORDS
prehensive computational analysis of the effects of MF and NSP on the Heat exchanger; metal
solidification process of RT82 paraffin (PCM matrix) in a large-scale shell- foam–PCM; nano-PCM;
and-tube latent thermal energy storage unit (of heat exchanger form). The phase change material
developed 2D transient model developed on Ansys Fluent 15.0 software (PCM); solidification process
was initially verified using available literature experimental data. The pro-
cess performance was tested for 5% Al2O3 nanoparticles and various MFs
[i.e. aluminum (Al), copper (Cu), nickel (Ni) and titanium (Ti)] with varied
porosity (96–100%). The computed mean and spatial temperature and sol-
idified degree of the PCM block showed a drastic acceleration of the solidi-
fication process with the b MF technique rather than with the
nanoparticles system. The solidification performance increased in the direc-
tion of MF-thermal conductivity increase, i.e. Cu > Al > Ni > Ti, and
material porosity decrease. These conditions allow rapid HTF heat recovery
and then stocking considerable thermal energy. However, the MF porosity
could not decrease below 95% to avoid a huge loss of material storage
(PCM), thereby diminishing the thermal storage capacity of the LTES unit.

1. Introduction
Phase change materials (PCMs) have been extensively investigated as passive thermal energy stor-
age in earlier decades [1]. They have several benefits, including low temperature decreases during
heat recovery, high latent heat, low vapor pressure, chemical inertness, stability and nontoxicity
[2–5]. The major challenge for PCMs is to increase energy storage due to the fluctuation of
resources such as solar energy, which is not accessible all day, and also to minimize energy con-
sumption [6]. PCMs possess a large latent heat, allowing it to store heat (charging cycle) before

CONTACT Atef Chibani chibaniatef@gmail.com Laboratory of Environmental Process Engineering, Department of


Chemical Engineering, Faculty of Process Engineering, University Salah Boubnider Constantine 3, Constantine, Algeria
ß 2023 Taylor & Francis Group, LLC
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 795

Nomenclature
A mushy zone constant Ø volume concentration of nano-particles
Cp specific heat capacity (kJ kg1 K1) q density (kg m3)
d diameter (m) e porosity
g gravitational acceleration (m s2)
L latent heat capacity (kJ kg1) Subscripts
P pressure (N m2)
T temperature (K) i initial
K permeability (m2) l liquefied phase
KB Boltzmann constant s solidified phase
t time (s) npcm nano-PCM
f liquid fraction of nano-PCM HTF heat transfer fluid
!
V velocity (m s1) np nano-particles
V volume (m3) pcm base paraffin
k thermal conductivity (W m1 K1) PCM phase change material
pc phase change
Greek ref reference
SENS sensible
b thermal expansion coefficient (K1) LAT latent
m dynamic viscosity (kg m1 s1)

releasing it (discharging cycle), both at narrow temperature interval [7]. The thermal conductivity
of commercialized PCMs is often poor (i.e. 0.1–0.7 W/mK for Paraffins [8]), which affects heat
exchange from the HTF (i.e. Heat transfer Fluid) to the PCM or the inverse [9]. Metal foams
[10–15], fins [16, 17] as well as nano-sized particle dispersion in the PCM body [18, 19] are
among the most recent solutions for overcoming for poor PCMs conductivity. Metal foams (MF)
as well fins, as compared to nano-sized particles (NSP) dispersion, not only greatly increase the
weight and expense of storage systems, but also lower the quantity of PCM that forms the body
of energy storage [18]. However, the correct choice of nanoparticle type is very important for
improving the process efficiency.
Several experimental studies focusing on analyzing PCMs storage system with MF and NSP
are available in the literature. Zhao et al. [20] examined the heat exchange of paraffin wax-PCM
in copper foams during the melting and solidification cycles. The use of MF boosts the global
heat transfer by 3–10 times during the melting phase, whereas the solidification time, on the
other hand, is cut in half. Li et al. [21] investigated heat transfer in paraffin–MF open-cell during
charging, where seven porous copper MFs were tested. Temperature distribution inside the sys-
tem were monitored and the impacts of foam morphological. factors on the wall. temperature
and temperature .uniformity within the foam body were explored. Ebadi et al. [22] evaluated the
charging mechanism of a thermal energy .storage device packed with PCM and metal wire mesh.
The impact of inserting copper wire mesh with two distinct porosities within a cylindrical LHTES
system exposed to three different isothermal surface temperatures has been studied. The impact
of adding metal wire mesh and isothermal surface temperatures on Nusselt number, heat flow
and energy stored was investigated. Zhang et al. [23] investigated the heat transfer properties of a
composite PCM made of paraffin and MF during the melting cycle. The scientists reported that a
paraffin/copper foam composite outperformed pure paraffin in terms of heat transfer perform-
ance. The distribution of temperature in the composite PCM–MF block was more uniform than
that of sole paraffin. In another hand, in order to verify the thermal performance of a nano-PCM
unit, the size, shape and material of the nanoparticles were examined by Jebasingh and Arasu
[24]. The outed results revealed that the thermal conductivity of the system increased gradually
while the latent heat decreased with nanoparticles dispersion. Bashar and Siddiqui [25] studied
the transient melting and heat exchange properties in a nano-PCM rectangular. The nanomaterial
796 A. CHIBANI ET AL.

tested were silver, Ag, CuO, Al2O3 and multiwall nanotubes. When compared to a pure PCM,
the inclusion of nanoparticles ameliorated the thermal performance of all hybrid nano-PCM com-
posites. Kumar et al. [26] studied the influence of various percentages (0.5–2.0%, w/w) of nano-
Si3N4 particles on paraffin thermal conductivity. The addition of nano-Si3N4 particles to basic
paraffin enhanced its thermal conductivity substantially; however, the relationship between the
thermal improvement and the nanoparticles dosage raised nonlinearity. Recently, using PCM as a
heat exchanger to absorb reaction heat in metal hydride H2-storage reactors provides an elegant
solution for effective thermal management [7, 27–34].
In parallel to the experimental directions, many computational analyses have been reported,
where different adopted models were resolved using Fluent CFD software. For example, Liu et al.
[12] developed a computational model to estimate the PCM melting process in a shell-and-tube
LHTES unit incorporating MF. The impact of HTF circumstances and the structural factors of
the porous system on the LHTES unit’s thermal properties are investigated. Xu et al. [11] quanti-
tatively evaluated the melting performance of PCMs in MF-based horizontal concentric-tube, con-
sidering natural convection. The findings revealed that the ideal filling. height ratio of porous
media is 0.7. Mesalhy et al. [35] numerically assessed the impact of MF on the thermal conductiv-
ity of PCMs. They noted that a decrease in MF-porosity could result in an increase of the melting
rate but damp the convection motion. In series of computational studies, Buonomo et al. [36–38]
evaluated the performance of many LHTES systems with PCM partially filled with aluminum
foam in local thermal equilibrium. The results showed that the heat transfer in the LHTESS was
enhanced when the metal foam thickness increased. The liquid fraction had a faster increase for
the initial time for the configuration with the lowest metal foam thickness but after this configur-
ation resulted, the slowest increase in liquid fraction was observed. Colla et al. [39] studied the
ano-PCMs system for enhancing energy storage and passive cooling applications. he results
showed that the Al2O3 nanoparticles penalizes the thermal conductivity while the addition of
Carbon Black nanoparticles leads to enhancement of 35% and 24%, for RT20/CB and RT25/CB,
respectively [39]. The Al2O3-based nano-PCMs were found to be unstable and a certain nanopar-
ticle deposition was observed; on the contrary, the CB based nano-PCMs were extremely stable
[39]. The MF-porosity and pore size effect MF on the thermal performance of PCMs unit has
been predicted by Sundarram and Li [40] using a three dimensional. finite element method. The
effects of pore size and porosity were more noticeable at high heat released and low convective
cooling circumstances. Elsayed [41] numerically studied how multi-nanoparticle mixes (Cu, Al,
TiO2, and SiO2) in a cylindrical container can recover the performance of solid-solid phase transi-
tion materials. The thermal comportment of the composites was demonstrated and described
under the impact of a nanoparticles concentration. More recently Mehdi’s research team has pub-
lished many interesting articles on PCMs solidification using different improvement techniques
[42–46]. The first configuration consists on applying multiple PCMs, cascaded MF and nano-
sized particles in the shell-and-tube latent energy storage system (LESS) for solidification
enhancement [44]. Compared to the module of one PCM with no cascaded foam or nanoparticles
the complete solidification time .saving was up to 94%, depending on the number of .cascaded
foam .segments and number of multiple. PCMs. The second design focused on using twisted .fins
for solidification enhancement in a triple-tube LESS [42]. The outcomes reveal that the operation.
of four twisted fins. decreased the solidification. time by 12.7% and 22.9%. compared with four
.straight fins and the .no-fins cases, respectively. The third configuration consists on using circu-
lar .fins with inline. and staggered arrangements. to enhance heat transfer in a triple tube LESS
[45]. The effect of number, position and temperature of the HTF Tubes on PCM solidification in
a multi-tube LTES system has been also studied by Li et al. [46]. The results indicates that the
specific .concern to the HTF tube. arrangement and .number should be made to .improve the dis-
charging .process attending free .convection impact in the phase change heat .storage.
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 797

However, despite the wide range of reactor designs used in PCM meting/solidification proc-
esses, there is a surprising paucity. of data on large-scale operation, with the majority of available
research focusing on lab- to pilot-scale operations. The development of industrial latent thermal
energy storage units necessitates the design and simulation of large-scale PCM-based solidification
processes. Furthermore, understanding the thermal behavior of PCMs, i.e. heat transfer, in large-
scale operations is critical, since the reactor design, the number and size of pipes, and their
arrangement all have a significant impact on the solidification process’ efficiency.
As a result, the primary goal of this research is to quantitatively study the solidification process
of paraffin-PCM embedded with various metal foams, i.e. of varying vide fraction, in a large-scale
shell-and-tube heat exchanger with a specified shape. To investigate the impacts of foam material,
four distinct scenarios are used (Cooper, Aluminum, Nickel and Titanium) on the paraffin’s tem-
perature and solidification kinetics. A series of simulations were also used to investigate the
impacts of MF-porosity. Simulated runs for an Al2O3 nano-PCM were also carried out to demon-
strate the benefits of the MF-PCM over pure and nano-PCMs in terms of heat transfer and
solidification rate acceleration.

2. Physical description of the unit


In the present study, the investigated latent heat storage unit is a horizontal shell-and-tube type
heat exchanger (diameter: 0.4 m) with multi-tubes (25 hexagonal tubes), where heat transfer fluid
(HTF) flows through the twenty-five inner hexagonal tubes. The outer side of the reactor shell is
adiabatic. The horizontal cross section of the cylindrical heat exchanger is shown in Figure 1a.
Gravity is considered in the vertical direction, toward the heat exchanger’s bottom. It should be
noted that the system is studied in a two-dimensional condition due to having no changes in the
y direction, with the assumption of considering a long length and ignoring the wall effects. The
annular space (between the hexagonal tubes and the thermal unit’s container) of the heat exchan-
ger is filled with either a pure phase change material (RT82), RT82-PCM loaded in various metal
foams (PCM-MF system) or (ii) Al2O3 nano-PCM (Al2O3 nanoparticles is homogeneously dis-
persed in the RT82-PCM). Both the PCM-MF and nano-PCM systems were adopted to enhance
the PCMs low thermal conductivity (k), thereby accelerating heat transfer during the energy stor-
age process. Al2O3 nanoparticles are selected among different available nanomaterials because of
their inertness and best thermal conductivity (i.e. 38.493 W/m K, against 0.2 W/m K for RT82).
The PCM solidification process was investigated in four distinct MFs (Cu, Al, Ni, and Ti), for the
same purpose. The thermophysical characteristics of the nanoparticles and the MFs selected are
listed in Table 1. In the numerical simulations, six scenarios were explored, designated by cases 1
to 6 in Table 2, in which the solidification process of pure paraffin (case 1) was compared to that
of nano-PCM using 5% Al2O3 (case 2) and MF-PCM utilizing the four mentioned metals at four
different porosities (case 2 to case 6).

3. Mathematical formulation
Thermal energy is transported by conduction in the PCM solid state and subsequently by both
conduction and natural convection once the liquid phase is generated in the PCM. Heat exchange
is enhanced in the porous-MF-PCM thanks to the high thermal conductivity of the MF material.
The porous structure inside the PCM was modeled using a thermal equilibrium model and an
enthalpy–porosity approach for the phase change effect. In the thermal equilibrium model, both
the PCM and the porous medium are kept at the same temperature [47, 48]. Bernardo et al. [36]
performed a thorough justification of the local thermal equilibrium in their numerical assess-
ments of LTES systems with aluminum foam. By using scale analysis, Bernardo et al. [36] calcu-
lated the magnitude of the local temperature difference between foam, solid matrix, and PCM in
798 A. CHIBANI ET AL.

Figure 1. Dimensional schematic of the shell-and-tube LTES system (a), and configuration of the mesh after grid independence
analysis (b).

Table 1. Properties of materials used in the present numerical study.


q (Kg m3) k (W m1 K–1) Ts (K) Cp (J Kg–1 K–1) L (J/Kg) l (Kg m1 s1)
ql ¼ 770/(0.001
Paraffin (T-358)þ1), qs ¼ 950 0.20 354 2000 176000 0.001  exp(-4.25 þ 1790/T)
Copper, Cu [48] 8933 350–401 1086 381 – –
Aluminum, Al [48] 2700 205–230 660 871 – –
Nickel, Ni [48] 8907 89 1455 460.6 – –
Titanium, Ti [48] 4850 7.44 – 544.25 – –
Aluminum oxide 3980 38.493 – 778 – –
(Al2O3), ɸ ¼ 20 nm
[8, 48, 50]

Table 2. The different investigated cases of the MF-PCM systems.


Case System Porosity
Case 1 Pure paraffin e ¼ 100%
Case 2 Paraffin þ 5% nano-particle Al2O3 e ¼ 100%
Case 3 Paraffin þ metal foam e ¼ 96%
Case 4 Paraffin þ metal foam e ¼ 97%
Case 5 Paraffin þ metal foam e ¼ 98%
Case 6 Paraffin þ metal foam e ¼ 99%
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 799

solid and liquid phases relative to the total temperature difference. The balance between the local
heat transfer at the metal and PCM interface and the inertia term for the metal component vali-
dated the thermal equilibrium technique [36]. Additionally, the term for the number of cells in a
square channel is (2n)2, and the proportion has an order of magnitude of n3, whereas in a cubic
geometry, the number of cells is of order n3, and the fraction has an order of n4 [36]. The local
thermal equilibrium is satisfactory if the number of pores per inch equals 20, so (20)4 [36]. The
local thermal equilibrium is presumable in light of this theory.
The following are the main assumptions of the present model [8, 47–49]:

 The liquid phase of the nano-PCM samples is laminar, noncompressible and Newtonian in
character, as is their mobility in shell container.
 The volumetric expansion or contraction of samples during phase transition in the shell con-
tainer is not taken into account.
 Shell container material is not considered to be in the computational domain due to its poor
heat conductivity and external border insulation.
 Due to the material tubing’s higher thermal conductivity, the thickness of the tubing is under-
estimated and the temperature variance of the HTF is neglected.
 The Boussinesq approximation is used to calculate density variance and normal convection-
driven buoyancy.
 The gravity acceleration is in the negative Y axis direction (as shown in Figure 1)

3.1. Governing equations


Voller’s enthalpy approach is the most extensively utilized to simulate phase transition at the
moment. For the phase transition of the pure PCM (without MF or nanoparticles), the model is
based on the following equations:

 Mass conservation [48]


!!
r: v ¼ 0 (1)
 Momentum conservation
The phase change is accounted for by the porosity model that treats the transition zone as a
porous zone with a liquid percentage of 0–1 [48]:
 !  
@v ! ! ! !   !
qpcm þ v r: v ¼ r P þ r lr:! v  qpcm !
g ð1  b:ðT  Tm ÞÞ þ S (2)
@t

In Eq. (2), the source term is defined as [48]:


! ð1  fl Þ2 !
S ¼ A!
v ¼ Amush 3 v (3)
fl þ 0:001
Amush refers to the consecutive number in the mushy region. This constant is often between 105
and 107 [49]. The approximation of Boussinesq is utilized, and Amush is taken as 105 [49].

 Energy equations [48]


 
@hsens ðTÞ ! dhlat ! !
q þ r:hsens ðT Þq v ¼ rðkrT Þ q þ V : r: hlat (4)
@t dt
800 A. CHIBANI ET AL.

The sensible heat hsens(T), i.e. heat stored in the form of temperature increase in the liquid or
the solid phase, is given as [19]:
ðT
ð Þ
hsens T ¼ href þ CP dT (5)
Tm

href is the reference enthalpy at Tref and Cp is the specific heat capacity at constant pressure. The
energy stored during PCM phase transition, hlat(T), is given as [19]:
hlat ðT Þ¼ f ðT Þ:DhSL (6)
where DHS–L is the total latent heat of the solid–liquid phase transition and ‘f’ is the liquid frac-
tion (0  f  1). f is given as [50]:
8
>
> T  TS
< 0T  T
s
f ¼ Ts  T  Tl (7)
>
> Tl  Ts
:1 TT l

subscripts s and l denote the PCM solid and liquid phases, respectively.

3.2. Nano-PCM thermo-physical properties relations


Nanoparticles will have varying degrees of impact on the nano-thermophysical PCM’s characteris-
tics. The thermophysical characteristics of nano-fluids and nano-PCMs may be modeled and cal-
culated using a variety of models that have been documented and are often used in the literature
today. For the nano-PCM employed in this study, the thermophysical property relationships
(Eqs. (8)–(14), for density, specific heat capacity, latent heat, and thermal expansion coefficient,
respectively) presented below are approximatively appropriate [8, 49, 51–53]. It should be noted
that in all these works [8, 49, 51–53], single phase assumption for the mixture of nanoparticles
and PCM has been adopted for the nano-PCM system.
qnpcm ¼ ð1ØÞqpcm þ Øqnp (8)
ð1ØÞqpcm cppcm þ Øqnp cpnp
cpnpcm ¼ (9)
qnpcm
ð1ØÞqpcm Lpcm
Lnpcm ¼ (10)
qnpcm
ð1ØÞqpcm bpcm þ Øqnp bnp
bnpcm ¼ (11)
qnpcm
where Ø denotes the volume percentage of nanoparticles, Cp is specific heat, and the subscripts
np, npcm, and pcm denote the nanoparticle, nano-PCM, and pure PCM, respectively. The effect-
ive thermal conductivity of the nano-PCM is provided by [54]:
  sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
knp þ 2kpcm  2 kpcm  knp Ø KB Tnpcm
knpcm ¼   kpcm þ 5:104 qpcm Øcppcm f ðT, ØÞ (12)
knp þ 2kpcm þ kpcm  knp Ø qnp dnp

where KB is Boltzmann constant, 1.381  1023 J K1, and dnp is the particle diameter. The cor-
rection factor f(T,Ø) is defined as [54]:
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 801

Tnpcm
f ðT, ØÞ ¼ ð2:8217  102 Ø þ 3:917  103 Þ þ ð3:0669  102 Ø  3:91123  103 Þ (13)
Tref
The effective dynamic viscosity is estimated as follow [55]:
lnpcm ¼ 0:983eð12:958:ØÞ :lpcm (14)
It is important to note that the combination of the nano-metric size of particles with their low
percentage in the PCM mass (typically 5%) could affect some physicochemical properties of the
system insignificantly (e.g. density, viscosity), but it can affect the thermal conductivity of the sys-
tem (research item).

3.3. Metal-foam
For the MF-PCM system, the formulation of the constitutive equations depends on the choice of
the unknowns and the assumptions made. By way of example, the three equations adopted in the
Fluent thermophysical simulation code are as follows:

 Mass conservation [47]:


!
r: V ¼ 0 (15)
 Momentum conservation [47]:

 In the x direction
  !
qpcm @u 1 !  @P lpcm 2 1  f l Þ2 l qCpcm juj !
þ r: V u ¼  þ ðr uÞ  Amush 3ð u  þ pffiffiffiffi V u
e @t e @x e fl þ 0:001 K K
(16)
 In the y direction
  !
qpcm @v 1 @P lpcm 2 1  fl Þ 2 l qCpcm jvj !
!
þ r:ð V vÞ ¼  þ ðr vÞ  Amush 3ð v þ pffiffiffiffi V v
e @t e @y e fl þ 0:001 K K
þ ðqbÞpcm g:ðT  Tm Þ
(17)
 Energy equations [47]:
h i @T !    @f
eðqCp Þpcm þ ð1  eÞðqCp Þmf þ qðqCp Þpcm V :r:T ¼ r keff rT  eðqLÞpcm : (18)
@t @t
In the thermal equilibrium model, volume average thermal conductivity is assumed between
the PCM and the porous medium as [36, 47]:
keff ¼ ð1  eÞkMF þ ekpcm (19)
The individual k (thermal conductivity) of the MFs and PCM are given in Table 1. e is varied
between 0.96 and 1. The liquid fraction (0  f  1) f is calculated using Eq. (7).

3.4. Numerical model


Fluent 15.0 software was used to solve the above model. The energy charging mode was numeric-
ally solved using finite volume discretization and computational fluid dynamics (CFD). A fixed-
802 A. CHIBANI ET AL.

grid computational domain is used to solve the governing differential equations [Eqs. (1)–(19)]
and their unique boundary conditions. To process the current pressure–velocity coupling, a
PRESSURE-based segregated solver and a Semi-Implicit pressure linked equation SIMPLE algo-
rithm are used. For analyzing the pressure correction, the QUICK differencing method and
PRESTO are used. Pressure, velocity, and thermal energy all had under-relaxation factors of 0.3,
0.7 and 1, respectively. For time discretization, a transitory formulation was used. Values of 107,
105 and 105 were used as convergence conditions for the energy, momentum, and continuity
equations, respectively.

3.5. Time step and grid independence tests


To ensure the accuracy of our model, time step and grid independency tests for the domain of
computation were examined. The paraffin solidification kinetics (in the reactor of Figure 1) when
aluminum is used as a MF with a porosity of 96% was simulated for (a) three time steps, 0.1, 0.5,
and 1 s, for a grid size of 12541 cells and (b) three grid size, 7326, 12541 and 18964 cells, for a
fixed time step of 1 s. The obtained simulation findings for both cases are plotted in Figures 2a
and 2b, respectively. As seen from both sub-figures, all solidification curves are perfectly superim-
posed, confirming the stability of the numerical model. As a result, a mean grid size of 7326 and
a time step of 1 s are chosen in all simulations.

4. Model verification
Our model is appropriate for simulating the solidification process of PCMs in LESE, as established in
this section through utilizing the experimental data of Al-Abidi et al. [5, 56] for validation. These scien-
tists assessed the solidification and melting of RT82 paraffin in a triplex-tube heat exchanger equipped
with internal/external fins. Figures 3a and 3b confront our simulation outcomes with those of Al-Abidi
et al. [5, 56], in terms of (a) average-PCM temperature and (b) average liquid fraction during the dis-
charging cycles. The experimental circumstances of Abidi et al. [5, 56] were done in the caption of
Figure 3. For both sub-figures, perfect predictions of the solidification parameters (i.e. PCM average
temperature and solidification kinetics) were obtained by our model. Overall, the mean absolute per-
centage errors between numerical (our results) and experimental results (of Al-Abidi et al.) varied
between 1 and 7% (throughout the simulation time, 3000s), as calculated by the mean deviation equa-
tion (jTexp–Tmodj/Texp). Therefore, our model can be used for analyzing thermal. storage operations dur-
ing solidification cycles with high accuracy level.

5. Results and discussion


5.1. Effect of different MF and porosity on heat transfer and solidification kinetics
In this section, the impact of four MFs (Al, Cu, Ni, and Ti) of varying porosity (from 100 to
96%) on the solidification process of RT82 paraffin in the heat exchanger of Figure 1 was assessed
and compared with pure and nano-PCM systems. The nano-PCM system utilized Al2O3 nanopar-
ticles at 5% (v/v). Al2O3 was chosen because it is the most often utilized nanomaterial in melting
and solidification [57]. Moreover, the sole dosage of 5% was selected according to several studies
revealing the best performances (on melting and solidification) of low nano-oxides loadings (5%)
in LTES [8, 49]. Also, using one concentration of nano-oxide could reduce the number of data to
be analyzed.
Figures 4a–4d show the influence of Al2O3 nanoparticles (5%) and different MFs (at various
porosities, i.e. 96–99%) on the average temperature profiles of the PCM matrix throughout the
solidification period (25,000 s). Table 2 lists the many studied cases (cases 1–6). The temperature
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 803

Figure 2. Liquid fraction-based time step (a) and mesh size (b) independency tests during paraffin solidification in the MF-PCM
system (MF: aluminum, porosity: 96%). All curves in (a) and (b) are perfectly superimposed.

of the HTF inside the channels is kept at 343 K at all times, whereas that of the MF-PCM system
is 366 K (the paraffin is liquid at t ¼ 0). Figures 5a–5d show the corresponding impacts of the six
cases of Table 2 on the PCM liquid fraction profiles.
First, as a common statement, the temperature of the PCM-bed, for the different cases, showed
two regimes: a sudden drop during the first stage of solidification (t < 1000s), followed by a
second one where an obvious quench of the temperature decrease is obtained for up to 25,000 s.
This trend is ascribed to the transfer mode, which is thermal convection between the MF (or the
wall of the cooled HTF) and the PCM (depending on cases). A rapid transfer is ensured during
the initial stage by the high conductive MFs, which accelerate the PCM cooling process. The
quench obtained at the send regime is caused by the solid layer formed between the MF and the
804 A. CHIBANI ET AL.

Figure 3. Model verification by the experimental results of Al-Abidi et al. [5, 56] [initial PCM temperature: 366 K, HTF
Temperature: 341 K, HTF flow rate: 8.3 L min1].

liquid PCM. This layer acts as thermal resistance, limiting the heat evacuation from the PCM
body (liquid part). The inverse of this phenomenon was observed early for the melting pro-
cess [48].
From Figures 4a–4d, it is seen that the order of temperature decrease is: Pure PCM < nano-
PCM < MF-PCM, and the lower the MF porosity, the lower the achieved temperature. For case 1,
the pure-PCM temperature is 351 (4% of loss) at 1000 s and 348 K at 25,000 s (5% of loss). For
this case, 2 and 42% of the PCM were solidified at t ¼ 1000 and 25,000 s, respectively. These
lower losses in the PCM temperature (and solid fractions) are due to the very low conductivity of
the RT82 paraffin (0.2 W/m K), which alters the release of its heat to the cooled HTF. Dispersing
nano-Al2O3 in paraffin (case 2) enhanced the solidification process; the temperature losses
became 4.2 and 5.2% at t ¼ 1000 and 25 000 s, respectively, with amelioration in the solid fraction
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 805

Figure 4. Variation of average bed temperature of the MF-PCM (and nano-PCM) systems over time for different MF materials
(Al, Cu, Ni, Ti) and porosities (i.e. denoted by cases 1–6, with the porosity characteristics are shown in Table 2).

achieved 10 and 65%, respectively. Thus, the resulted impact of nano-particles dispersion was
more pronounced on the solid fraction (20% increase is recorded at t ¼ 25 000 s, as compared to
the pure PCM), which is in good agreement with the outcomes of Chibani et al. [8] for the RT82
melting in the presence of 5% Al2O3. This is owing to the nano-particles-induced enhancement
of the thermal conductivity of the paraffin, with accelerating the release of the PCM energy to
the HTF. Experimental evidence of the nano-PCM strategy for LTES unit has been provided by
Harikrishnan et al. [6] who studied the comportment of a hybrid nano-PCM experimentally with
paraffin wax and CuO/TiO2 nanoparticles at 1% (w/w). They compared the thermal conductivity,
viscosity and stability of the pure PCM and the nano-PCM systems. The conclusion was that the
nanoparticles prevent paraffin decomposition, and consequently, the nano-PCM matrix could be
more stable. Moreover, the heating/cooling processes of the nano-PCM system are faster than
paraffin alone; the melting. time decreased by 29.8% and the cooling. time with 27.7%.
On the other hand, the MF-PCM tests resulted in rapid reduction of the PCM temperature
and liquid fraction of the PCM. For a progressive reduction of e (MF porosity) form 100% (case
1, pure paraffin) to 99% (case 6), 98% (case 5), 97% (case 4) and 96% (case 3), mean losses in the
PCM temperature (as compared to case 1) of 4.5% for Al-MF, 5% for Cu-MF and 4.6% for Ni-
MF were recorded at t ¼ 1000s. However, higher solid fractions were recorded at this instant, as
listed below:
806 A. CHIBANI ET AL.

Figure 5. Variation of average PCM liquid fraction over time for different MF materials (Al, Cu, Ni, Ti) and porosities (i.e. denoted
by cases 1 to 6, with the porosity characteristics are shown in Table 2).

 Cu-MF: 40, 50, 60 and 70% for e ¼99, 98, 97 and 96% (case 6 to 3 gradually, Figure 5b),
 Al-MF: 28, 40, 46 and 52% for e ¼99, 98, 97 and 96% (case 6 to 3 gradually, Figure 5a),
 Ni-MF: 20, 27, 32 and 40% for e ¼99, 98, 97 and 96% (case 6 to 3 gradually, Figure 5c),
 Ti-MF: 5 to 10% for e ¼99 to 96% (case 6 to 3 gradually, Figure 5d).

The MF-PCM strategy is more performant at advanced discharging time. For example, at
t ¼ 5000s, the solid fraction data increased to:

 Cu-MF: 67, 85, 96 and 100% for e ¼ 99, 98, 97 and 96% (case 6 to 3 gradually, Figure 5b),
 Al-MF: 60, 75, 87 and 85% for e ¼ 99, 98, 97 and 96% (case 6 to 3 gradually, Figure 5a),
 Ni-MF: 44, 57, 66 and 78% for e ¼ 99, 98, 97 and 96% (case 6 to 3 gradually, Figure 5c),
 Ti-MF: 22 to 30% for e ¼ 99–96% (case 6 to 3 gradually, Figure 5d).
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 807

Thus, the solidification process in metallic foams happens extremely fast, and the temperature
of the MF-PCM matrix drops very quickly. Moreover, efficient heat exchange could be ensured
with porosity decreasing, but not unlimitedly, as this could reduce thermal storage capacity. In
fact, operating with lower porosity means lower mass of PCM, which is the latent heat storage
matrix [20, 22, 23, 48, 57, 58]. Mesalhy et al. [35] have treated numerically the enhancement of
the thermal conductivity of a PCM using MF. Their results showed that the porosity reduction
increased the melting efficiency but dampen. the convection motion. and the storage. capacity.
The minimum temperatures that can be achieved for cases 1 and 2 are 348.5 and 347.5 K,
respectively, at t ¼ 25 000 s (Figure 4). For the metallic PCM foams, the HTF temperature (343
K) can be reached rapidly at much reduced times, i.e. 4000, 5500, 8000 and 16,000 s for e(Cu-
MF) ¼ 96, 97, 98 and 99%, respectively (Figure 4b). For Al-MF, the PCM reached the HTF tem-
perature at 8000, 10,000 and 15,000 s for e ¼ 96, 97 and 98% (Figure 4a). However, much
extended time is required for cooling the PCM for the cases of Ni and Ti-metal foams
(Figures 4c and 4d). These results are in line with the data in Figure 6, demonstrating the time
required to reach maximum solidification percentage (indicated above barres) for the different
MF-PCM systems in Table 2. As seen, the highest solidified potions were achieved with a shorter
time for higher porous Cu- and the Al-MF. The benefits of Cu- and Al-MFs are linked to their
greater thermal conductivity than Ti and Ni, as seen in Table 1. In fact, the higher the MF porous
solid system’s thermal conductivity, the greater the heat transmission rates. It should be noted
that the insignificant impact of Ti-MF on heat transfer (Figure 4d) and solid fraction (Figure 5d)
was ascribed to the loo low thermal conductivity of Ti (7.44 W/m K) against that of Ni (89
W/m K).

5.2. Spatial distribution of the solid fraction during the solidification process
The contour diagrams of the PCM liquid/solid fraction after 1 h of solidification are shown in
Figure 7 for the six investigated cases (i.e. pure PCM, nano-PCM, and MF-PCM with different e).
Due to the increased heat transfer rate due to PCM incorporation in porous metal foams, the
MF-PCM systems generated a large solidified zone, dependent on the porosity and the type of
MF. Heat loss by convection is the dominant mechanism at the earlier stage of the PCM solidifi-
cation; however, after the formation of a thick solid layer around the HTF tubes, both convection
(from the liquid PCM to the solid part) and conduction (in the solid layer of the PCM) modes
co-produced. According to Figure 7, the thickness of the PCM-solid layer increased remarkably
with the porosity decrease (from 100% to 96%). Correspondingly, Cu- and Al-metals ensured the
PCM’s higher solidification rate due to their excellent thermal conductivity. For the pure PCM,

Figure 6. Variation of time required to reach maximum solidification percentage (indicated above barres), for the different MF-
PCM systems of Figure 5 (cases 3–6 of Table 2).
808 A. CHIBANI ET AL.

Figure 7. Distribution of solid (blue)/liquid (red) PCM-fractions for cases 1 to 6 during the solidification process of the PCM,
Nano-PCM and MF-PCMs, at 3600 s.

only low quantity of paraffin was solidified (the shell adjacent to the channels side). The thickness
of the solid layer increased by about 10% with Al2O3 dispersion in the PCM mass. For the case
of Ti-MF, the PCM-solid shell formation rate is comparable to that of the nano-PCM system,
which is mainly caused by the low thermal conductivity of Ti (as compared to the other investi-
gated nanomaterial). Complete solidification of the paraffin was reached for Cu-MF at e ¼ 96%
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 809

Figure 8. Temperature contours field for all the cases during the solidification process of the and MF-PCMs, at 3600 s (Al, Cu,
Ni, Ti).
810 A. CHIBANI ET AL.

Figure 8. Continued.

(Figure 7). As has been assumed that the Al2O3-paraffin is perfectly mixed, the nano-PCM system
could compared with the other MF system based on thermal conductivity (for the same volume
fraction). Based on the values of thermal conductivity provides in Table 1, the thermal
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 811

conductivity of Al2O3 (38.493 W/m K) is higher than that of Ti (7.44 W/m K) and lower than
that of Al (205–230 W/m K), Cu (350–401 W/m K) and Ni (89 W/m K). Thus, the heat transfer
in the PCM matrix could be better in Al2O3-paraffin Nano-PCM compared to Ti-MF, but lesser
than the other MFs (Al, Cu and Ni). Therefore, the solid fractions of in Nano-PCM system could
be greater than Ti but less than Al-, Cu-, and Ni-based MF-PCM systems (see Figure 7).
In Figures 8a–8d, the spatial distribution of the MF-PCM temperature at 3600s is given for
the various metal foams (Al, Cu, Ni et Ti) at different porosities (96–99%). Note that because of
the lower performance of nano-PCM and the pure paraffin), compared to metal foams systems,
the spatial distribution of the temperature (Figure 8) was given only for the MFs systems.
According to the solidification distribution of Figure 7, the reactor cooling begins from the HTF
tubes where the temperature is minimal (343 K). The cooled zones around the HTF tubes
become more extended for porosity decreases from 99 to 96%. This is related to the fact that
decreasing e could increase the mass of the MF, thereby increasing the heat transfer. This obser-
vation is maintained for the different adopted MF systems, as seen in Figures 8a–8d. Another
important comment is that the whole temperature distribution followed the order obtained for
the average data provided early in Figures 4a–4d in terms of process sensitivity to the type of the
MF. The higher the conductivity of the used MF, the higher the spatial heat recuperated from the
heated PCM. Therefore, Cu- and Al-based MF favors more evacuation of heat from the PCM
matrix, ensuring the best solicitation performance in the whole body of the unit. This is clearly
shown in Figures 8a and 8b, where big areas of the PCM-MF system achieve the HTF tempera-
ture at only 3600s, particularly for the case of e ¼ 96% cases. It should be also mentioned that the
temperature on the upper sides of the Ti cases (Figure 8d) is relativley higher than the other
regions, which is not observed in the other three cases (Al, Cu and Ni). A systematic explanation
was not found for this phenomenon, but it should be noted that the difference between the hot
region (in red) and that in yellow in note considerable (4 K), so in reality, there is no significant
temperature distribution in the case of Ti, due to its lower thermal conductivity! In conclusion,
this section’s outcomes confirm that the use of Cu or Al is healthier for attaining higher thermal
performance during the solidification process of paraffin PCMs.

6. Conclusion
In summary, the solidification process of paraffin in a large-scale thermal energy storage unit
with porous MF happens extremely fast. The temperature of the MF-PCM matrix and the liquid
fraction drops very quickly. Moreover, efficient heat exchange could be ensured with porosity
decreasing, but not unlimitedly, as this could decrease the thermal storage capacity. the order of
temperature decrease is: Pure PCM < nano-PCM < MF-PCM. Besides, the solidification per-
formance increased with MF-thermal conductivity increase, i.e. Cu > Al > Ni > Ti. These find-
ings were obtained by determining the average and spatial distribution of temperature and liquid
fraction over solidification for different investigated systems. The results of this study may pro-
vide practical guidelines for developing more effective LTES units for large-scale applications. It
is also critical to include cost estimates for all simulated scenarios, including the costs associated
with the inclusion of nanoparticles and various metal foams, in order to better gauge its usability
in large-scale practical applications. However. The cost estimation requires knowing the reactor
length (in addition to the section) to calculate the reactor volume (mass of the PCM). This means
that all simulation results (for the different scenarios) could be re-conducted in 3D reactor form,
which is another big project. Thus, because our article deals with only 2D simulation (horizontal
section), an economical estimation of the presses is not possible. A future project will address all
these issues.
812 A. CHIBANI ET AL.

Acknowledgments
This research. was funded by Algeria’s Ministry of Education. and Science Research, as well as the General.
Directorate for Scientific. Research and Technological. Development (GD-SRTD) (project No.
A16N01UN250320220002).

ORCID
Atef Chibani http://orcid.org/0000-0002-5861-7498

References
[1] S. N. Nyamsi, I. Tolj, and M. Lototskyy, “Metal hydride beds-phase change materials: dual mode thermal
energy storage for medium-high temperature industrial waste heat recovery,” Energies, vol. 12, no. 20, pp.
3949, 2019. DOI: 10.3390/en12203949.
[2] R. Pakrouh, M. J. Hosseini, A. A. Ranjbar, and R. Bahrampoury, “A numerical method for PCM-based pin
fin heat sinks optimization,” Energy Convers. Manage., vol. 103, pp. 542–552, 2015. DOI: 10.1016/j.encon-
man.2015.07.003.
[3] C. Ji, Z. Qin, Z. Low, S. Dubey, F. H. Choo, and F. Duan, “Non-uniform heat transfer suppression to
enhance PCM melting by angled fins,” Appl. Therm. Eng., vol. 129, pp. 269–279, 2018. DOI: 10.1016/j.
applthermaleng.2017.10.030.
[4] C. Pan, S. Hoenig, C. H. Chen, S. Neti, C. Romero, and N. Vermaak, “Efficient modeling of phase change
material solidification with multidimensional fins,” Int. J. Heat Mass Transfer., vol. 115, pp. 897–909, 2017.
DOI: 10.1016/j.ijheatmasstransfer.2017.07.120.
[5] A. A. Al-Abidi, S. Mat, K. Sopian, M. Y. Sulaiman, and A. T. Mohammad, “Numerical study of PCM
solidification in a triplex tube heat exchanger with internal and external fins,” Int. J. Heat Mass Transfer.,
vol. 61, no. 1, pp. 684–695, 2013. DOI: 10.1016/j.ijheatmasstransfer.2013.02.030.
[6] S. Harikrishnan, K. Deepak, and S. Kalaiselvam, “Thermal energy storage behavior of composite using
hybrid nanomaterials as PCM for solar heating systems,” J. Therm. Anal. Calorim., vol. 115, no. 2, pp.
1563–1571, 2014. DOI: 10.1007/s10973-013-3472-x.
[7] H. El Mghari, J. Huot, L. Tong, and J. Xiao, “Selection of phase change materials, metal foams and geome-
tries for improving metal hydride performance,” Int. J. Hydrogen Energy, vol. 45, no. 29, pp. 14922–14939,
2020. DOI: 10.1016/j.ijhydene.2020.03.226.
[8] A. Chibani and S. Merouani, “Acceleration of heat transfer and melting rate of a phase change material by
nanoparticles addition at low,” Int. J. Thermophys., vol. 42, no. 5, pp. 1–16, 2021. DOI: 10.1007/s10765-
021-02822-z.
[9] M. M. Farid, A. M. Khudhair, S. Ali, and K. Razack, “A review on phase change energy storage: materials
and applications,” Energy Convers. Manage., vol. 45, nos. 9–10, pp. 1597–1615, 2004. DOI: 10.1016/j.encon-
man.2003.09.015.
[10] Z. Liu, Y. Yao, and H. Wu, “Numerical modeling for solid–liquid phase change phenomena in porous
media: shell-and-tube type latent heat thermal energy storage,” Appl. Energy, vol. 112, pp. 1222–1232, 2013.
DOI: 10.1016/j.apenergy.2013.02.022.
[11] A. J. Chamkha, A. Doostanidezfuli, E. Izadpanahi, and M. Ghalambaz, “Phase-change heat transfer of sin-
gle/hybrid nanoparticles-enhanced phase-change materials over a heated horizontal cylinder confined in a
square cavity,” Adv. Powder Technol., vol. 28, no. 2, pp. 385–397, 2017. DOI: 10.1016/j.apt.2016.10.009.
[12] Y. Xu, Q. Ren, Z. Zheng, and Y. He, “Evaluation and optimization of melting performance for a latent heat
thermal energy storage unit partially filled with porous media,” Appl. Energy, vol. 193, pp. 84–95, 2017.
DOI: 10.1016/j.apenergy.2017.02.019.
[13] V. Joshi and M. K. Rathod, “Thermal transport augmentation in latent heat thermal energy storage system
by partially filled metal foam : a novel con figuration,” J. Energy Storage, vol. 22, December 2018, pp. 270–
282, 2019. DOI: 10.1016/j.est.2019.02.019.
[14] M. Ghalambaz and J. Zhang, “Conjugate solid–liquid phase change heat transfer in heatsink filled with
phase change material-metal foam,” Int. J. Heat Mass Transfer, vol. 146, pp. 118832, 2020. DOI: 10.1016/j.
ijheatmasstransfer.2019.118832.
[15] M. Ghalambaz, S. M. Hashem Zadeh, S. A. M. Mehryan, I. Pop, and D. Wen, “Analysis of melting behavior
of PCMs in a cavity subject to a non-uniform magnetic field using a moving grid technique,” Appl. Math.
Model., vol. 77, pp. 1936–1953, 2020. DOI: 10.1016/j.apm.2019.09.015.
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 813

[16] A. Singh, M. P. Maiya, and S. S. Murthy, “Effects of heat exchanger design on the performance of a solid
state hydrogen storage device,” Int. J. Hydrogen Energy, vol. 40, no. 31, pp. 9733–9746, 2015. DOI: 10.1016/
j.ijhydene.2015.06.015.
[17] S. N. Nyamsi, F. Yang, and Z. Zhang, “An optimization study on the finned tube heat exchanger used in
hydride hydrogen storage system: analytical method and numerical simulation,” Int. J. Hydrogen Energy,
vol. 37, no. 21, pp. 16078–16092, 2012. DOI: 10.1016/j.ijhydene.2012.08.074.
[18] M. Gorzin, M. J. Hosseini, M. Rahimi, and R. Bahrampoury, “Nano-enhancement of phase change material
in a shell and multi-PCM-tube heat exchanger,” J. Energy Storage, vol. 22, no. June 2018, pp. 88–97, 2019.
DOI: 10.1016/j.est.2018.12.023.
[19] B. Kok, “Examining effects of special heat transfer fins designed for the melting process of PCM and nano-
PCM,” Appl. Therm. Eng., vol. 170, pp. 114989, 2020. DOI: 10.1016/j.applthermaleng.2020.114989.
[20] C. Y. Zhao, W. Lu, and Y. Tian, “Heat transfer enhancement for thermal energy storage using metal foams
embedded within phase change materials (PCMs),” Solar Energy, vol. 84, no. 8, pp. 1402–1412, 2010. DOI:
10.1016/j.solener.2010.04.022.
[21] W. Q. Li, Z. G. Qu, Y. L. He, and W. Q. Tao, “Experimental and numerical studies on melting phase
change heat transfer in open-cell metallic foams filled with paraffin,” Appl. Therm. Eng., ” vol. 37, pp. 1–9,
2012. DOI: 10.1016/j.applthermaleng.2011.11.001.
[22] S. Ebadi, S. H. Tasnim, A. A. Aliabadi, and S. Mahmud, “An experimental investigation of the charging
process of thermal energy storage system filled with PCM and metal wire mesh,” Appl. Therm. Eng., vol.
174, no. April, pp. 115266, 2020. DOI: 10.1016/j.applthermaleng.2020.115266.
[23] P. Zhang, Z. N. Meng, H. Zhu, Y. L. Wang and S. P. Peng, “Melting heat transfer characteristics of a com-
posite phase change material fabricated by paraffin and metal foam q,” Appl. Energy, vol. 185, pp. 1971–
1983, 2017. DOI: 10.1016/j.apenergy.2015.10.075.
[24] E. J. B. V. A. A., “A comprehensive review latent heat thermal conductivity nanoparticle dispersed phase
change material low-temperature applications,” Energy Storage Mater., vol. 24, no. May 2019, pp. 52–74,
2020. DOI: 10.1016/j.ensm.2019.07.031.
[25] M. Bashar and K. Siddiqui, “Experimental investigation of transient melting and heat transfer behavior of
nanoparticle-enriched PCM in a rectangular enclosure,” J. Energy Storage, vol. 18, no. March, pp. 485–497,
2018. DOI: 10.1016/j.est.2018.06.006.
[26] P. M. Kumar, R. Anandkumar, K. Mylsamy and K. B. Prakash, “Materials today: Proceedings experimental
investigation on thermal conductivity of nanoparticle dispersed paraffin (NDP),” Materials Today:
Proceedings, 2020. DOI: 10.1016/j.matpr.2020.02.798.
[27] Y. Ye, Y. Yue, J. Lu, J. Ding, W. Wang, and J. Yan, “Enhanced hydrogen storage of a LaNi5 based reactor
by using phase change materials,” Renew. Energy, vol. 180, pp. 734–743, 2021. DOI: 10.1016/j.renene.2021.
08.118.
[28] T. Alqahtani, S. Mellouli, A. Bamasag, F. Askri, and P. E. Phelan, “Thermal performance analysis of a metal
hydride reactor encircled by a phase change material sandwich bed,” Int. J. Hydrogen Energy, vol. 45, no.
43, pp. 23076–23092, 2020. DOI: 10.1016/j.ijhydene.2020.06.126.
[29] T. Alqahtani, A. Bamasag, S. Mellouli, F. Askri, and P. E. Phelan, “Cyclic behaviors of a novel design of a
metal hydride reactor encircled by cascaded phase change materials,” Int. J. Hydrogen Energy, vol. 45, no.
56, pp. 32285–32297, 2020. DOI: 10.1016/j.ijhydene.2020.08.280.
[30] L. Tong, Y. Yuan, T. Yang, P. Benard, C. Yuan, and J. Xiao, “Hydrogen release from a metal hydride tank
with phase change material jacket and coiled-tube heat exchanger,” Int. J. Hydrogen Energy, vol. 46, no. 63,
pp. 32135–32148, 2021. DOI: 10.1016/j.ijhydene.2021.06.230.
[31] H. Q. Nguyen and B. Shabani, “Thermal management of metal hydride hydrogen storage using phase
change materials for standalone solar hydrogen systems: an energy/exergy investigation,” Int. J. Hydrogen
Energy, vol. 47, no. 3, pp. 1735–1751, 2022. DOI: 10.1016/j.ijhydene.2021.10.129.
[32] J. Yao, et al., “A continuous hydrogen absorption/desorption model for metal hydride reactor coupled with
PCM as heat management and its application in the fuel cell power system,” Int. J. Hydrogen Energy, vol.
45, no. 52, pp. 28087–28099, 2020. DOI: 10.1016/j.ijhydene.2020.05.089.
[33] L. Tong, J. Xiao, P. Benard and R. Chahine, “Thermal management of metal hydride hydrogen storage res-
ervoir using phase change materials,” Int. J. Hydrogen Energy, vol. 44, no. 38, pp. 21055–21066, 2019. DOI:
10.1016/j.ijhydene.2019.03.127.
[34] H. El Mghari, J. Huot, J. Xiao, H. El, J. Huot and J. Xiao, “Analysis of hydrogen storage performance of
metal hydride reactor with phase change materials,” Int. J. Hydrogen Energy, vol. 44, no. 54, pp. 28893–
28908, 2019. DOI: 10.1016/j.ijhydene.2019.09.090.
[35] O. Mesalhy, K. Lafdi, A. Elgafy and K. Bowman, “Numerical study for enhancing the thermal conductivity
of phase change material (PCM) storage using high thermal conductivity porous matrix,” Energy Convers.
Manage., vol. 46, no. 6, pp. 847–867, 2005. DOI: 10.1016/j.enconman.2004.06.010.
814 A. CHIBANI ET AL.

[36] B. Buonomo, H. Celik, D. Ercole, O. Manca and M. Mobedi, “Numerical study on latent thermal energy
storage systems with aluminum foam in local thermal equilibrium,” Appl. Therm. Eng., vol. 159, no. June,
pp. 113980, 2019. DOI: 10.1016/j.applthermaleng.2019.113980.
[37] B. Buonomo, O. Manca, S. Nardini and R. E. Plomitallo, “Numerical study on latent heat thermal energy
storage system with PCM partially filled with aluminum foam in local thermal equilibrium,” Renew. Energy,
vol. 195, pp. 1368–1380, 2022. DOI: 10.1016/j.renene.2022.06.122.
[38] B. Buonomo, O. Manca, S. Nardini and R. E. Plomitallo, “Numerical investigation on shell and tube latent
thermal energy storage partially filled with metal foam and corrugated internal tube,” IJHT, vol. 40, no. 4,
pp. 895–900, 2022. DOI: 10.1115/HT2022-81806.
[39] L. Colla, L. Fedele, S. Mancin, L. Danza, and O. Manca, “Nano-PCMs for enhanced energy storage and pas-
sive cooling applications,” Appl. Therm. Eng., vol. 110, pp. 584–589, 2017. DOI: 10.1016/j.applthermaleng.
2016.03.161.
[40] S. S. Sundarram and W. Li, “The effect of pore size and porosity on thermal management performance of
phase change material infiltrated microcellular metal foams,” Appl. Therm. Eng., vol. 64, nos. 1–2, pp. 147–
154, 2014. DOI: 10.1016/j.applthermaleng.2013.11.072.
[41] A. O. Elsayed, “Numerical study on performance enhancement of solid–solid phase change materials by
using multi-nanoparticles mixtures,” J. Energy Storage, vol. 4, pp. 106–112, 2015. DOI: 10.1016/j.est.2015.09.
008.
[42] X. Sun, J. M. Mahdi, H. I. Mohammed, H. S. Majdi, W. Zixiong, and P. Talebizadehsardari, “Solidification
enhancement in a triple-tube latent heat energy storage system using twisted fins,” Energies, vol. 14, no. 21,
pp. 7179, 2021. DOI: 10.3390/en14217179.
[43] M. E. Tiji, et al., “Natural convection effect on solidification enhancement in a multi-tube latent heat stor-
age system: effect of tubes’ arrangement,” Energies, vol. 14, no. 22, pp. 7489, 2021. DOI: 10.3390/
en14227489.
[44] J. M. Mahdi, H. I. Mohammed, E. T. Hashim, P. Talebizadehsardari and E. C. Nsofor, “Solidification
enhancement with multiple PCMs, cascaded metal foam and nanoparticles in the shell-and-tube energy
storage system,” Appl. Energy, vol. 257, no. June 2019, pp. 113993, 2020. DOI: 10.1016/j.apenergy.2019.
113993.
[45] X. Sun, et al., “Investigation of heat transfer enhancement in a triple tube latent heat storage system using
circular fins with inline and staggered arrangements,” Nanomaterials, vol. 11, no. 10, pp. 2647, 2021. DOI:
10.3390/nano11102647.
[46] M. Li, et al., “Solidification enhancement in a multi-tube latent heat storage system for efficient and eco-
nomical production: effect of number, position and temperature of the tubes,” Nanomaterials, vol. 11, no.
12, pp. 3211, 2021. DOI: 10.3390/nano11123211.
[47] A. Chibani, S. Merouani and C. Bougriou, “The performance of hydrogen desorption from a metal hydride
with heat supply by a phase change material incorporated in porous media (metal foam): Heat and mass
transfer assessment,” J. Energy Storage, vol. 51, no. September 2021, pp. 104449, 2022. DOI: 10.1016/j.est.
2022.104449.
[48] A. Chibani, S. Merouani, and F. Benmoussa, “Computational analysis of the melting process of Phase
change material-metal foam-based latent thermal energy storage unit: the heat exchanger configuration,” J.
Energy Storage, vol. 42, no. April, pp. 103071, 2021. DOI: 10.1016/j.est.2021.103071.
[49] A. Chibani, et al., “A strategy for enhancing heat transfer in phase change material-based latent thermal
energy storage unit via nano-oxides addition: A study applied to a shell-and-tube heat exchanger,” J.
Environ. Chem. Eng., vol. 9, no. 6, pp. 106744, 2021. DOI: 10.1016/j.jece.2021.106744.
[50] Z. Khan, Z. A. Khan, and P. Sewell, “Heat transfer evaluation of metal oxides based nano-PCMs for latent
heat storage system application,” Int. J. Heat Mass Transfer, vol. 144, pp. 118619, 2019. DOI: 10.1016/j.
ijheatmasstransfer.2019.118619.
[51] S. Ebadi, S. H. Tasnim, A. A. Aliabadi, and S. Mahmud, “Melting of nano-PCM inside a cylindrical thermal
energy storage system: numerical study with experimental verification,” Energy Convers. Manage., vol. 166,
no. March, pp. 241–259, 2018. DOI: 10.1016/j.enconman.2018.04.016.
[52] A. Chibani, S. Merouani, C. Bougriou, and A. Dehane, “Heat and mass transfer characteristics of charging
in a metal hydride-phase change material reactor with nano oxide additives: the large,” Appl. Therm. Eng.,
vol. 213, no. February, pp. 118622, 2022. DOI: 10.1016/j.applthermaleng.2022.118622.
[53] J. M. Mahdi and E. C. Nsofor, “Solidification of a PCM with nanoparticles in triplex-tube thermal energy
storage system,” Appl. Therm. Eng., vol. 108, pp. 596–604, 2016. DOI: 10.1016/j.applthermaleng.2016.07.
130.
[54] R. S. Vajjha and D. K. Das, “Experimental determination of thermal conductivity of three nanofluids and
development of new correlations,” Int. J. Heat Mass Transfer, vol. 52, nos. 21–22, pp. 4675–4682, 2009.
DOI: 10.1016/j.ijheatmasstransfer.2009.06.027.
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 815

[55] I. Sarani, S. Payan, S. A. Nada and A. Payan, “Numerical investigation of an innovative discontinuous dis-
tribution of fins for solidification rate enhancement in PCM with and without nanoparticles,” Appl. Therm.
Eng., vol. 176, no. December 2019, pp. 115017, 2020. DOI: 10.1016/j.applthermaleng.2020.115017.
[56] A. A. Al-Abidi, S. Mat, K. Sopian, M. Y. Sulaiman, and A. T. Mohammad, “Experimental study of melting
and solidification of PCM in a triplex tube heat exchanger with fins,” Energy Build., vol. 68, pp. 33–41,
2014. DOI: 10.1016/j.enbuild.2013.09.007.
[57] P. T. Sardari, H. I. Mohammed, D. Giddings, G. S. Walker, M. Gillott, and D. Grant, “Numerical study of
a multiple-segment metal foam–PCM latent heat storage unit: effect of porosity, pore density and location
of heat source,” Energy, vol. 189, pp. 116108, 2019. DOI: 10.1016/j.energy.2019.116108.
[58] D. Zhou and C. Y. Zhao, “Experimental investigations on heat transfer in phase change materials (PCMs)
embedded in porous materials,” Appl. Therm. Eng., vol. 31, no. 5, pp. 970–977, 2011. DOI: 10.1016/j.applth-
ermaleng.2010.11.022.

You might also like