Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

Numerical Heat Transfer, Part B: Fundamentals

An International Journal of Computation and Methodology

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

Effects of different heat transfer enhancers in


PCM-based latent thermal energy storage for
hydrogen storage in activated carbon bottle

Atef Chibani, Slimane Merouani & Aissa Dehane

To cite this article: Atef Chibani, Slimane Merouani & Aissa Dehane (27 Nov 2023): Effects of
different heat transfer enhancers in PCM-based latent thermal energy storage for hydrogen
storage in activated carbon bottle, Numerical Heat Transfer, Part B: Fundamentals, DOI:
10.1080/10407790.2023.2286254

To link to this article: https://doi.org/10.1080/10407790.2023.2286254

Published online: 27 Nov 2023.

Submit your article to this journal

Article views: 76

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
https://doi.org/10.1080/10407790.2023.2286254

Effects of different heat transfer enhancers in PCM-based


latent thermal energy storage for hydrogen storage in
activated carbon bottle
Atef Chibania , Slimane Merouanib, and Aissa Dehaneb
a
Research Center in Industrial Technologies CRTI, Algiers, Algeria; bLaboratory of Environmental Process
Engineering, Department of Chemical Engineering, Faculty of Process Engineering, University Salah Boubnider
Constantine 3, Constantine, Algeria

ABSTRACT ARTICLE HISTORY


This study presents a numerical simulation of a packed bed reactor filled Received 24 April 2023
with activated carbon for hydrogen storage, incorporating phase change Revised 10 November 2023
material (PCM) with various heat transfer enhancers. The simulations were Accepted 15 November 2023
performed using the Fluent ANSYS. The main goal is to explore the use of
KEYWORDS
PCM as a heat manager (reactor coolant) and optimize its performance by Activated carbon bed;
investigating the effects of various enhancers, including nanoparticles (NP), hydrogen storage; mass
metal foam (MF), and fins, on the melting efficiency of PCM in activated transfer enhancers;
carbon beds. Outcomes showed that the hydrogen adsorption capacity of numerical simulation; Phase
25 mmol/g was found to be independent of the type of PCM enhancer change Material (PCM)
used in the system. The results showed that the MF provided the best
cooling performance of the activated carbon bed, with a surface heat flux
of 2750 W/m2 and a surface heat transfer coefficient of 410 W/m2K, while
fins provided the highest melting rate of the PCM. In the presence of fins,
metal foam, and nanoparticles, the fusion rate of PCM has been enhanced
by 64, 32, and 8%, respectively, compared to pure PCM. These findings
provide valuable insights into the design of packed bed reactors for hydro­
gen storage and can be used to optimize the thermal performance of the
system. Based on the obtained findings, it is recommended that future
studies on H2 storage in solid beds should be focused on the optimization
of fins, metal foams, use of other nanoparticles, integration of composite
PCMs, adoption of chemically modified or nanoporous carbon beds, and
possibly the employment of hybrid enhancers.

1. Introduction
Hydrogen has been recognized as a promising energy carrier due to its high energy content and
potential to replace fossil fuels [1]. One kilogram of hydrogen contains about 141.8 MJ of energy,
which is more than double the energy content of gasoline (about 46.4 MJ/kg) and natural gas
(about 55.5 MJ/kg) [1]. It can be produced from various sources such as renewable energy, nat­
ural gas, and biomass through different methods including steam methane reforming, water elec­
trolysis, and biomass gasification [1–6]. However, hydrogen storage remains a major challenge as
it requires high-pressure tanks or low-temperature storage in liquid or solid forms [7]. These
storage methods have limitations such as high cost, safety concerns, and low storage capacity [8].
Therefore, the development of efficient and cost-effective hydrogen storage systems is crucial for
the widespread adoption of hydrogen as an alternative energy source [9].

CONTACT Atef Chibani chibaniatef@gmail.com Research Center in Industrial Technologies CRTI, P.O.Box 64, Cherag,
Algiers 16014, Algeria.
� 2023 Taylor & Francis Group, LLC
2 A. CHIBANI ET AL.

Nomenclature
−1 −1
CP specific heat, J.kg .K q Density, kg m−3
D diameter, m a Thermal diffusivity of the liquid PCM,
DH molar enthalpy of reaction at standard m2s−1
conditions, J.mol−1
DS molar entropy of reaction under standard Subscripts
conditions, J.mol−1.K−1
E activation energy, J mol−1 i Initial
h convection heat transfer coefficient, l Liquefied phase
W.m−2.K−1 s Solidified phase
M molar mass of hydrogen, kg.mol−1 f fluid
P hydrogen pressure, Pa g gas
T temperature, K m melting
t time, s w wall
R universal gas constant, J.mol−1.K−1 b bed
S source term from reaction, kg.m−3.s−1 eff effective
~
v Superficial velocity vector, m s−1 pcm Base paraffine
v kinematic viscosity (m2 s−1) ref Reference
V velocity, m.s−1 PCM Phase change material
MF Metal foam
Greek NP Nanoparticles
AC Activated Carbon
b thermal expansion coefficient (K-1) sens sensible
m Dynamic viscosity (kg m-1 s-1) lat latent
e Metal foam porosity

Hydrogen storage on solid supports such as metal hydrides has been extensively studied in
recent years as an alternative to high-pressure or cryogenic storage methods [10]. Various mate­
rials, such as rare earth alloys, intermetallic compounds, and metal-organic frameworks, have
been investigated for their potential as hydrogen storage media [11]. One promising candidate
is magnesium hydride, which has a high hydrogen storage capacity of 7.6 wt% and can release
hydrogen at moderate temperatures [12]. Other metal hydrides, such as LiH, NaAlH4, and
LiBH4, have also been investigated for their hydrogen storage properties [13]. With the combin­
ation of PCM to the different stocking systems, Ben-Maad et al. [14] quantitatively investigated
the use of LaNi5 metal hydride (MH) with LiNO3-3H2O PCM for managing the heat generated
during hydrogen absorption. They found that 80.0% of the stored hydrogen in the MH could
be discharged using only the heat stored in the PCM during the charging cycle. Additionally,
researchers have explored the use of carbon nanotube (CNT) nanoparticles to improve heat
transfer in PCM-based systems for building ventilation applications [15]. Other studies have
demonstrated that the conductivity of the PCM can be improved by incorporating heat transfer
fluid (HTF) tubes or metal foams, which can significantly enhance the performance of the MH-
PCM storage system [16–18]. Mghari et al. [19] reported that the MH-PCM reactor had a
favorable impact on the dehydriding/hydriding period, with the latent heat of the PCM having
a greater effect on the system’s reaction than thermal conductivity. It is worth mentioning that
metal hydrides suffer from several drawbacks, including limited reversibility and high operating
temperatures, which make them unsuitable for many practical applications [8]. To overcome
these limitations, new materials and novel approaches, such as catalytic enhancement and nano­
structuring, have been proposed to improve the hydrogen storage properties of metal hydrides
[8]. Despite these efforts, solid state hydrogen storage remains a challenging and active research
area in the field of hydrogen energy.
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 3

Hydrogen storage on activated carbon has gained significant attention due to its high surface
area, low cost, and stability. Activated carbon is a porous material with a large surface area, typic­
ally ranging from 500 to 2500 m2/g, which provides high adsorption capacity and fast adsorption
kinetics [20]. Hydrogen adsorption on activated carbon is known to be an exothermic process
during the adsorption step and an endothermic one during the desorption step. This reversible
behavior makes activated carbon a promising candidate for hydrogen storage. Many studies have
been conducted on hydrogen storage on activated carbon, and several methods have been pro­
posed to enhance hydrogen uptake and release. In their study, Hirscher and Panella [21] deter­
mined that activated carbon with a BET surface area of 2560 m2/g had a maximum storage
capacity of 4.5 wt% at a temperature of 77.4 K. On the other hand, Kojima et al. [22] investigated
the adsorption and desorption of hydrogen in various carbon materials at two different tempera­
tures, namely 296 K and 77 K. Their findings showed that superactivated carbon with a specific
surface area (SSA) of 3220 m2/g could adsorb 1.3 wt% of hydrogen at 296 K and 5 wt% of hydro­
gen at 77 K, surpassing the values of conventional activated carbon, graphite nanofiber, and sin­
gle-walled carbon nanotube. Besides, numerical simulations are increasingly being employed to
gain insights into hydrogen storage in activated carbon. For example, Paggiaro et al. [23] devel­
oped a two-dimensional model to investigate the thermal effects during the high-pressure filling
of a cryo-adsorptive hydrogen storage tank using powdered activated carbon as the adsorbent.
The simulation results showed good agreement with experimental data. Momen et al. [24]
employed a numerical model based on the solution of the 2D transport equations for mass,
momentum, and energy in porous media to study hydrogen charging in packed bed storage
tanks. Their findings revealed that high thermal conductivity packed beds could overcome cap­
acity limitations caused by thermal heating. Xiao et al. [25] created a finite element model to
simulate the heat and mass transfer in an activated carbon hydrogen storage tank, considering
the heat capacity of the adsorbed phase, contact thermal resistance between the AC bed and the
steel wall, and the inertial resistance of high-speed charging hydrogen gas. In addition to the
aforementioned works, other theoretical studies [26,27] have been conducted to enhance and
accelerate the improvement of hydrogen storage in solid phase.
To make the process practical, an efficient heat transfer system is required to recover the heat
released during the adsorption step and supply the heat needed during desorption. Many studies
have investigated the use of heat transfer fluids (HTF) and heat exchangers [28–33] to improve
the energy efficiency of the process [34–38]. In recent years, efforts have been made to optimize
the heat recovery process in hydrogen storage systems by using different heat transfer fluids
(HTFs) such as oils or molten salts and nanofluids [39–41]. However, the use of phase change
materials (PCMs) has been proposed as an alternative to HTFs due to their high energy density
and ease of use [42–45]. PCMs are able to store and release large amounts of thermal energy dur­
ing phase change, which makes them suitable for use in heat recovery systems for hydrogen stor­
age on activated carbon. Compared to the sensible thermal energy storage in materials like
concrete and water, the latent heat storage unit may save up to 90% of mass and area to store the
same amount of thermal energy [46]. Generally, phase change materials (PCMs) with large latent
heat, high thermal conductivity, adequate melting temperature, chemical stability, low cost, non­
toxic and non-corrosive are in high demand for an efficient application in thermal storage units
[47]. Research in this area has shown promising results, with PCMs being able to significantly
improve the thermal performance of the system and reduce the overall energy consumption. In
the theoretical study of Chibani et al [48], a qualitative and quantitative study was conducted to
investigate the performance of hydrogen adsorption (on activated carbon bed) in the presence of
PCM (during charging and dormancy periods) and oil (over the discharge phase). Taking into
account the enthalpy contours and liquid fraction of PCM (18 and 30% at 400 and 4000s,
respectively), it was concluded that the thermal conductivities of the employed PCM and acti­
vated carbon bed should be enhanced for a better performance of the hydrogen stocking system.
4 A. CHIBANI ET AL.

In another study [49] an industrial scale for hydrogen storage was investigated in the presence of
metal foam (e ¼ 0.8–1)-PCM system. The heat recovery during the hydrogen adsorption was
remarkably enhanced with the incorporation of metal foam (copper). Additionally, with the
decrease in MF porosity (from 1 to 0.8), a substantial improvement (several times) was retrieved
for the heat flux, heat transfer coefficient, and melting rate of PCM.
Despite the growing interest in using PCMs as thermal energy storage in hydrogen storage sys­
tems, some scarcity is observed for the use of PCM in the case of adsorptive storage tanks.
Additionally, there is a lack of comparative studies on different heat transfer enhancement techni­
ques for large-scale reactors. While some studies have investigated the use of metal foam, fins or
nanoparticles as enhancers, few have compared their effectiveness or explored their potential in
combination with other techniques. This gap in knowledge highlights the need for further
research to optimize heat transfer in PCM-enhanced hydrogen storage systems.
Therefore, the objective of this study is to compare the performance of different PCM heat
transfer enhancers, including metal foam, fins and nanoparticles, for use in large-scale hydrogen
storage reactors. By investigating the effectiveness of these enhancers in enhancing heat transfer,
we aim to provide insight into the optimization of PCM-enhanced hydrogen storage systems.
This study will contribute to filling the gap in knowledge regarding the comparative effectiveness
of various heat transfer enhancement techniques.

2. Physical and numerical model


The computational domain (meshed surface) was symmetrical with respect to the vertical axis.
Therefore, only the right-half of the domain was simulated. The schematic diagram of the sug­
gested activated carbon hydrogen storage bottle (cylindrical) is shown in Figure 1(a). The steel
bottle height (H) is 150 cm (1.50 m), with internal (R1) and external (R2) radii of 11.5 and
18.0 cm, respectively (Figure 1(a)). The feeding entrance of the reactor is 1.5 cm in radius and
10 cm in height (expansion volume in Figure 1(a)). In this system, the AX-21 of activated carbon

Figure 1. Schematic of the physical model (a), and computational domain and grid structure for AC-nano-PCM, AC-PCM with
MF and AC-PCM with fins (b).
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 5

Table 1. Material properties [27,62–66].


Activated carbon Specific heat capacity (J/kg K) 825
Density (kg/m3) 517.6
Thermal conductivity (W/m K) 0.646
Porosity 0.490
Hydrogen Density Ideal gas
Thermal conductivity (W/m K) 0.206
Specific heat capacity (J/kg K) 14,700
Viscosity (Pa s) 8.411 � 10–6
Viscous resistance coefficient (m-2) 8.29055 � 107
Inertial resistance coefficient (m-1) 7586
Steel wall Density (kg/m3) 7830
Specific heat capacity (J/kg K) 468
Thermal conductivity (W/m K) 13
RT 22 HC (PCM) Density (kg/m3) liquid 700
solid 760
Specific heat capacity (J/kg K) 2100
Thermal conductivity (W/m K) 0.210
Viscosity (Pa s) 0.020
Phase change temperature (K) 293.15–296.15
Latent heat of fusion (kJ/kg) 190
Copper (used in metal foam/fines) Density (kg/m3) 8978
Specific heat capacity (J/kg K) 381
Thermal conductivity k (W/m K) 387.6
Nanoparticles (CuO) Particles diameter, dp (nm) 15–30
Density(kg/m3) 6500
Thermal conductivity (W/m K) 17.991
Specific heat capacity(J/kg K) 536

(AC) is packed in the bottle’s interne (of R1 ¼ 11.5 cm), whereas the PCM-based system (i.e.
pure, with nanoparticles, with MF, or with fins) was loaded in the annular section (11.5 cm
<R < 18 cm) (Figure 1(a)). Rectangular fins (thickness: 0.1 mm, length: 5.5 cm) was adopted.
Copper (Cu) is selected as the matter of foam while CuO was chosen as nanoparticles. This
choice is not arbitrary, it was because of their (Cu and CuO) high thermal conductivity as well as
their frequent use as enhancers in different thermal energy storage systems [50–52]. It is worth
mentioning that the thermal conductivity of nanoparticles is considered as the main parameter
affecting the thermal performance of the different thermal units. However, this is not the only
key parameter (i.e. thermal conductivity) controlling the selection of nanoparticles, because other
criteria, such as the impacts of nanoparticles on the different parameters of the Nano-PCM sys­
tem, namely: physicochemical stabilities, cost-effectiveness, specific heat, latent heat of fusion,
density, and viscosity, play a potent role in the adoption of these nanoparticles [53]. Furthermore,
the homogeneous emulsification of the nanoparticles and PCM is considered as a crucial process
for maximizing the thermal performance (reduction of melting/solidification rates of PCM and
increasing the energy storage capacity) of the Nano-PCM system [53]. Thermal–physical parame­
ters of copper and CuO in addition to those of the activated carbon, hydrogen, phase change
material (RT 22 HC), and the steel wall are shown in Table 1. It should be noted that for the
case of fins, the upper section of the reactor (PCM side) is slightly modified (Figure 1(b)) just to
include the last fin and achieve 5% of fins in the PCM volume. This modification has absolutely
no effect on the simulation results due to the negligible volume of PCM added (inferior to 0.1%).
For this study, a single percentage of 5% (v/v) of copper and CuO was selected as the optimal
dosage of the enhancer (nanoparticles, metal foam, or fins) in the PCM volume for each system.
It should be stressed here that this percentage (5%, v/v) has been shown to be effective in various
applications, such as paraffin wax and hydrated salt systems, and has been found to be an accept­
able choice for an optimum balance between the positive (increasing the thermal conductivity of
PCM) and negative (reducing the available latent heat of PCM, increasing the PCM’s viscosity
and possibly reducing the stability of Nano-PCM due to the sedimentation and agglomeration)
6 A. CHIBANI ET AL.

effects resulting from the addition of nanoparticles to PCM body [11,54–60]. In these studies, the
authors investigated the effect of varying percentages of the enhancer on the thermal performance
of PCM and found that 5% was an appropriate dosage for achieving significant enhancements in
heat transfer. Furthermore, Chibani et al. [61] found that increasing the percentage of nanopar­
ticles beyond 5% did not significantly enhance the thermal performance of PCM melting, and
may even decrease its stability. These findings suggest that 5% is an appropriate dosage of the
enhancer in the PCM volume to achieve significant enhancements in heat transfer while main­
taining the stability of the PCM. It is important to note that higher enhancer quantities can
reduce the amount of PCM available for storing energy, which can reduce the overall energy stor­
age capacity of the system. In addition, nanoparticles have a tendency to agglomerate at higher
dosage, thus increasing the viscosity of the PCM, which can have a detrimental effect on the ther­
mal performance of the system [55,59,61]. However, it is important to note that the optimal per­
centage may vary depending on the specific application and should be carefully determined
through experimentation. In this study, the selected percentage of 5% was deemed appropriate
for enhancing heat transfer while maintaining the stability of the PCM.

2.1. CFD model


2.1.1. For hydrogen storage
The main foundation of the CFD model commonly resides in mass., momentum and energy-con­
servation formulas of the hydrogen system. Actually, the following formulas strongly rely on
hydrogen, activated carbon and a steel tank wall. It is noteworthy to say that the adsorbed sub-
models adopted in this work involve elements of Dubinin-Asatkhov (D-A) adsorption-isotherms
and kinetics, which is applicable up to the critical point of hydrogen

� Mass conservation equation


In porous media, the mass conservation equation for hydrogen is given as [27,67]:
@ ðeb q g Þ @n
vÞ¼
þr:ðqg~ ð1 eb ÞqP MH2 (1)
@t @t
where n (mol kg−1) is the absolute adsorbed amount per unit of activated carbon, eb the bed
porosity of the activated carbon, MH 2 (kg mol−1) molecular_mass of hydrogen, qg (kg m−3)
the density of gaseous hydrogen, ~ v (m s−1) the superficial_velocity vector (~
v ¼ eb~
u , where ~
u
is the physical velocity) in porous media, qp (kg m−3) the Particle density of activated carbon.
� Momentum conservation
In the case of standard fluid flow, the momentum equation is stated as [68]:
@
vÞþr:ðqg~
ðq ~ v~
vÞ¼ rpþr: %s þqg~
g (2)
@t g
In porous media case, a momentum source term ~ S should be placed on the right side of the
above equation. The parameter s is the stress tensor and ~ S (N m−3) is the momentum source
term, which is based on the Ergun equation and includes viscous and inertial losses. Its com­
ponent in the “i” direction, “Si”, for simple homogeneous porous media, may be written as
[20]:
� �
l 1
Si ¼ v jvi
vi þC2 qg j~ (3)
k 2
The parameter l (Pa s) is the dynamic viscosity of hydrogen gas, 1/k (m−2) is the viscous
resistance coefficient, C2 (m−1) is the inertial resistance coefficient, j~
v j is the magnitude_of the
velocity vector and vi is the velocity component along the “i” direction.
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 7

� Energy conservation equation


The energy equation for the hydrogen storage tank represents the balance between the amount
of energy collected in the tank and the energy changes owing to convective flow, pressure
work, conductive and thermal dispersion fluxes, and heat release due to the adsorption pro­
cess. It may be given as [69]:
@� � X
vðqg Eg þ pÞ�¼r:ðkeff rT
eb qg Eg þð1 eb Þqp Es þr:½~ hi~
J i þ %s :~
vÞþSe (4)
@t i

The variable Se is an energy source parameter which is expressed as:


Se ¼ DHSm =MH2 (5)
The parameter DH (KJ/Kmol) represents the constant_isosteric heat of adsorption. The vari­
ation of DH is mainly due to the experimental circumstances as well as the structure of acti­
vated carbon. The energy source term (Se, W/m3) is related to the mass source term (Sm, kg
m−3 s−1) and represents the heat’s amount created per unit volume. Overall, for the mixing
features of the porous medium, the FLUENTTM software examines just two components. The
specific heat is expressed as the volumetric average of the gas and solid phases:
ðqcp Þm ¼ eb qg cpg þ ð1 − eb Þqp�cps (6)

�cps ¼ cps þqa cpa � cps þqa cpg (7)


�cps is the modified specific heat for the solid phase (adsorbent and adsorbate), where the influ­
ence of the hydrogen-adsorbed phase is included.
� Adsorption models
To determine the link between hydrogen adsorption density and the pressure and temperature
in the tank, the modified Dubinin-Astakhov(D-A) adsorption model is used. The absolute
adsorption isotherm is given by [68]:
" � �m � �#
� RT m p0
n ¼nmax exp ln (8)
aþ bT p

where, nmax (mol kg−1) is the limit adsorption, the value of nmax is 71.6 mol kg−1, m is a num­
ber generally close to a small integer and equal to 2 for most activated carbons. R is the uni­
versal gas constant, and p is the equilibrium pressure. The value of the saturation pressure p0
is 1470 MPa. The enthalpy a and the entropic factors b are 3080 J mol−1 and 18.9 J mol−1
K−1, respectively. The adsorption kinetics in the model is approximated by the linear driving
force (LDF) model [62]:
dn
¼ k ðn� − nÞ (9)
dt
The parameter n� is the adsorbed gas in equilibrium with the gas phase and k is a mass trans­
fer coefficient at an aggregated level. In this paper, a value of k ¼ 0.45 s−1 is assumed. In add­
ition, we select 10,500 J/mol as the constantisosteric heat of adsorption [62].

2.1.2. For the PCM system


Voller’s enthalpy-porosity formulation provided by the software is used for simulating paraffin
phase change. This model simplifies the phase change problem by providing the following bene­
fits: (i) the governing equations are equivalent to the equation for a unique phase, (ii) absence of
conditions to meet the interface and (iii) the enthalpy formulation declares a mixing zone
between phases, permitting us to follow the melting-front easily.
8 A. CHIBANI ET AL.

� Mass conservation [61]:


~ v¼ 0
r:~ (10)
� Momentum conservation [60]:
� �
v
@~ ~ vÞ ¼
qpcm v r:~
þð~ r~
Pþrðlr:~ g ð1 b:ðT Tm ÞÞþ~
v Þ qpcm~ S (11)
@t
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 to 1. In Equation 11, the source term is defined as
[57]:

~ ð1 f l Þ2
S¼ A~
v¼ Amush v
~ (12)
f 3l þ0:001
where Amush refers to the consecutive number in the mushy region. This constant is often
between 105 and 107 [55]. The approximation of Boussinesq is utilized, and Amush is taken as 105
[11,55,56].

� Energy equations [70]:

Enthalpie:
� �
@hsens ðTÞ dhlat ~ ~
q þr:hsens ðTÞq~
v¼rðkrTÞ q þV:r:hlat (13)
@t dt
Total energy:
@ � � � �X � �
~ e~
eqf Ef þð1 eÞqs ES þr: ~ keff rT
v ðqf Ef þPÞ ¼r: ~ hi ji þð%s :~
v Þ þSe (14)
@t i

p v2
With: E ¼ h þ (15)
q 2
The total energy per unit mass (E) contained in the system is comprised of three parts:
internal, kinetic and potential. The sensible-heat hsens(T), i.e. heat stored in the form of tempera­
ture increase in the liquid or the solid phase, is given as [70]:
ðT
ð Þ
hsens T ¼href þ CP dT (16)
Tm

href is the reference enthalpy at Tm ¼ 295 K and Cp is the specific heat capacity at constant pres­
sure. The energy stored during PCM phase transition, hlat(T), is given as [69]:
hlat ðTÞ¼ f ðTÞ:DhS L (17)
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 [71]:
8
>
> 0 T� TS
>
<T T
s
f¼ Ts � T� Tl (18)
>
> Tl Ts
>
: 1 T�Tl

Subscripts "s" and "l" denote the PCM solid and liquid phases, respectively.
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 9

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 thermo-physical simulation code are as follows:

� Mass conservation [56]:


r:~
V¼ 0 (19)
� Momentum conservation [56]:

- In the x-direction
� �
qpcm @u 1
ð~
þ r: VuÞ
e @t e
! (20)
@P lpcm 2 ð1 f l Þ2 l qCpcm juj ~
¼ þ ðr uÞ Amush 3 u þ pffiffiffi V u
@x e f l þ0:001 K K
- In the y-direction
� �
qpcm @v 1
~
þ r:ðVvÞ
e @t e
! (21)
@P lpcm 2 ð1 f l Þ2 l qCpcm jvj ~
¼ þ ðr vÞ Amush 3 v þ pffiffiffi V vþðqbÞpcm g:ðT Tm Þ
@y e f l þ0:001 K K

� Energy equations [56]:


h i @T @f
eðqCp Þpcm þð1 eÞðqCp Þmf þqðqCp Þpcm ð~
V:r:TÞ¼rðkeff rTÞ eðqLÞpcm : (22)
@t @t
In the thermal equilibrium model, volume average thermal conductivity is assumed between the
PCM and the porous medium as [56]:
keff ¼ð1 eÞkMF þekpcm (23)
For the nano-PCM, the resulting system was treated as a single system (perfectly mixed). The
equation driving this system, as well as the equation providing the thermos-physical properties
(density, specific heat capacity, latent heat, and thermal expansion coefficient, respectively) of the
nano-PCM system, are available in detail in our previous works [55,59,61].
The initial system’s (AC þ PCMs) temperature is 293.15 K. The charging time was fixed at
500 s. The outside of our tank is adiabatically insulated (Q ¼ 0). The hydrogen velocity at the bot­
tle entrance is fixed at 1 m/s. Conduction/conduction heat continuity between AC and reactor
wall:
n¼ 0
ðkeff rTAC kw rTw Þ:~ (24)
Conduction/conduction heat continuity between PCM and reactor wall/fin
n¼ 0
ðkpcm rTpcm kw rTw Þ:~ (25)

n¼ 0
ðkpcm rTpcm kfin rTfin Þ:~ (26)

n¼ 0
ðkpcm rTpcm kMF rTMF Þ:~ (27)
10 A. CHIBANI ET AL.

� The adiabatic (axis-symmetry) boundary condition:


rTAC ¼ rTpcm ¼ 0 (28)
The energy charging mode was numerically solved using finite volume discretization and compu­
tational fluid dynamics (CFD). A fixed-grid computational domain is used to solve the governing
differential equations (Eqs. (1)–(28)) and their unique boundary conditions. To process the cur­
rent pressure–velocity coupling, a PRESSURE-based segregated solver and a Semi-Implicit pres­
sure linked equation SIMPLE algorithm 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 10−7, 10−5, and 10−5 were used as convergence conditions for
the energy, momentum, and continuity equations, respectively.
To ensure the accuracy of the simulation results, the independence of the computational
domain’s time-step and grid resolution was evaluated. The charging process of hydrogen in the
activated carbon-MF-PCM system was simulated using four different grid resolutions (cell num­
bers) of 17364, 9561, 7585, and 6388 cells with three time steps of 0.5, 0.75, 1, and 1.5 s. The
average temperature profiles of the activated carbon bed were analyzed after three and four time
steps (with 9561 cells). The results revealed that the bed temperature increased significantly dur­
ing the initial loading phase (0 to 500 s) due to the exothermal adsorption of H2 on the activated
carbon. This temperature reached a maximum (387 K) at 500 s before being quenched and grad­
ually returning to the initial temperature after a longer time of 10000 s. The maximum bed tem­
perature was found to have an insignificant deviation of 5% maximum between the four different
cell numbers (17364, 9561, 7585, and 6388), indicating the strong stability of the numerical solu­
tion. Additionally, the temperature profiles remained stable for time steps of 0.75, 1, and 1.5 s,
with a maximum deviation of about 3–5% in the maximum bed temperature recorded. Therefore,
to increase the accuracy of the results and shorten computing time, a grid size of 9561 and a
time-step of 1s were used throughout the remaining sections.

3. Model validation
Model validation is a crucial step in numerical simulation studies as it ensures the accuracy and
reliability of the results obtained from the simulation. By comparing the simulation results with
experimental data, the validity of the mathematical model and numerical methods used can be

Figure 2. Model validation using the Hermosilla-Lara et al. [26] experimental data.
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 11

verified. This helps to increase confidence in the results obtained from the simulation and
improve the understanding of the physical phenomena involved. Therefore, in this section, our
model will be validated with experimental data of Hermo-silla-Lara et al. [20] before the results
and discussion section. The Hermosilla-Lara et al.’s experimental work was conducted using a
packed bed of activated carbon (cylindrical form), and the hydrogen adsorption was carried out
at different temperatures and pressures [20]. The authors measured the amount of hydrogen
adsorbed by the activated carbon as well as the heat of adsorption during the charging process.
The temperature evolution is several points in the AC bed was monitored. These profiles were
used as probes in our model validation section. The comparison between the experimental evolu­
tions and the simulated profiles (using our model) is shown in Figure 2.
The comparison between the experimental and simulated profiles showed excellent concord­
ance (Figure 2), with the model accurately capturing the experimental temperature evolution over
a time of 600 s. The temperature increased during the adsorption cycle, reaching a maximum at
200 s, and then decreased during the cooling phase. The model provided the following maximum
temperatures: 347, 340, and 310 K for points 3, 5, and 6, respectively, while the corresponding
experimental temperatures were 343, 334, and 312 K. This gives deviations of 1.2 (point 3), 1.8
(point 5), and −0.64% (point 6) for our simulations results compared to the experimental ones
(at 200 s). At 600 s, the experimental temperatures were 317, 309, and 300 K for points 3, 5, and
6, respectively, whereas the model predicted 318, 312, and 301 K. Therefore, at 600 s, our simula­
tion results are deviated by 0.32 (point 3), 0.97 (point 5), and 0.33% (point 6) compared to the
experimental ones. As a result, the deviations between the model and experimental temperatures
ranged from −0.64% to 1.8%, which are within acceptable limits for a numerical model. Overall,
the excellent agreement between the experimental and simulated profiles confirms the accuracy of
our model in predicting the thermal behavior of hydrogen adsorption on activated carbon bed.

4. Results
4.1. Hydrogen adsorption in the presence of mass transfer enhancers
With 5% of each enhancer, we simulated the average temperature and hydrogen concentration in
the activated carbon bed. The latent thermal unit (inducing PCM þ enhancers) surrounded the
activated carbon bed to absorb the generated heat from the exothermal reaction. The average
temperature and hydrogen concentration in the activated carbon bed were followed as a function
of time for the four systems, namely: PCM pure (no enhancer), nano-PCM, PCM with metal
foam, and PCM with fins. The results are shown in Figure 3(a,b).
The activated carbon (AC) temperature in all cases increased suddenly from 295.15 K at t ¼ 0
to 388 K at t ¼ 500 s, and then decreased near-exponentially up to t ¼ 10,000 s, Figure 3(a).
However, the AC temperature profile after 500 s was slightly different, where the PCM with metal
foam provided the best cooling action (lower temperature profiles after 500 s). The temperature
profile of the AC with PCM with fin provides slightly better cooling than the pure PCM, with a
difference of about 2 to 3%. The AC bed temperature with the nano-PCM exhibits the same
trend as that of the pure PCM. The sudden increase in AC temperature is independent of the
enhancer type and is due to the exothermic reaction that occurs quickly in the activated carbon
bed during the hydrogen adsorption process. The PCM with the MF system enhances the cooling
performance by increasing the heat transfer coefficient, as reported in the literature [17,72,73].
The MF provides a large interfacial area between the PCM and the activated carbon bed, leading
to efficient heat transfer [74]. Furthermore, the PCM with the MF system exhibits better thermal
conductivity than the other systems, leading to enhanced heat dissipation. In contrast, the PCM
with fins system provides a lower interfacial area (compared to MF), leading to a lower heat
transfer rate and reduced cooling performance. The PCM with nano-oxide additives shows lower
12 A. CHIBANI ET AL.

Figure 3. Evolution of (a) the average bed (AC) temperature and (b) the hydrogen concentration (adsorption capacity) with time
during the bottle charging using 1 m/s of hydrogen rate at the bottle entrance (PH2,inp. ¼ 0.048 MPa).

thermal conductivity, leading to lower heat dissipation and similar cooling performance to the
pure PCM system. This later observation (nano-PCM) is similar to that reported in [60]. On the
other hand, the adsorbed H2 concentration in the activated carbon bed increases linearly with
time for all systems, reaching a plateau at around 25 mmol/g (Figure 3(b)). A similar profile has
been recently reported by Chibani et al. [11,48,59,63,75] for hydrogen storage in metal hydride
system coupled with different nano-PCMs and PCM-MFs systems, for lab- and large-scale reac­
tors using AC or metal hydride. The similar H2 concentration profiles for all systems suggest that
the presence of enhancers in the PCM does not significantly affect the adsorption capacity of the
activated carbon bed. This trend results mainly from the quick exothermic adsorption of hydro­
gen on AC, as revealed in Figure 3(a) where the AC temperature increased suddenly and inde­
pendently of the latent thermal system. However, it is worth noting that higher quantities of
enhancers could potentially reduce the PCM mass, which serves as the storage site for the gener­
ated heat and hence reduces the amount of energy that can be stored. Additionally, the use of
nanoparticles as enhancers can lead to the detrimental phenomenon of nanoparticle agglomer­
ation and an increase in viscosity, which can negatively affect the performance of the system [55].
These factors should be considered when selecting the type and quantity of enhancers for PCM-
based hydrogen storage systems.
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 13

Providing specific references that support our findings about the order of superiority between
metal foam, fins, and nanoparticles for latent thermal energy storage is not possible because most
studies didn’t compare different enhancer systems, but compare the performance of the same

Figure 4. The melt fraction of the PCM (a), the overall heat transfer coefficient (b), and overall total heat flux (c) over time for
pure PCM, nano-PCM, PCM with MF and PCM with fins. 5% v/v for each enhancer.
14 A. CHIBANI ET AL.

enhancer at varying material types. However, some studies made comparisons between two
enhancers instead. In a series of recent publications on metal hydride-enhancer systems, Chibani
et al. [11,55,59,75] highlighted (i) the absence of a thermal efficiency difference between pure
PCM (paraffin) and nano-PCM [59] (ii) the huge improvement of the thermal energy storage by
using MF [11,75], and (iii) the dependence of the nanoparticles enhancing effect to the reactor
design (lab- or large scale) and the operating conditions [55].

4.2. Thermal behavior of the PCM-enhancer systems


In this section, the thermal behavior of PCM-enhancer systems was analyzed during hydrogen
charging in activated carbon. The melt fraction of the PCM, overall heat transfer coefficient, and
overall total heat flux were simulated over time for pure PCM and PCM with enhancers, includ­
ing nanoparticles, metal foam, and fins. The following is a detailed report of the findings. The
founding results are shown in Figure 4(a–c).
As seen, the PCM melting fraction increased linearly with time until 400 s, after which the
slope decreased progressively, leading to a near-plateau shape after 6000s, Figure 4(a). During
the rapid adsorption phase (<400 s), approximately the same fusion rate of PCM is obtained in
the presence of metal foam and fins. Thus, in these systems (PCM-foam and PCM-Fins), the
melting rate of PCM is enhanced by � two times compared to the cases of Nano-PCM and
pure PCM. Similarly, to the outcomes of Figure 3(a), in Figure 4(a) and at t > 500s, a clear dis­
crepancy is observed between the investigated systems (Foam-PCM, Nano-PCM, Fins-PCM,
and pure PCM) in terms of liquid fraction of PCM. Therefore, in comparison to the pure
PCM, its melting rate is improved by 64, 32, and 8% in the presence of fins, metal foam, and
nanoparticles, respectively. Consequently, the melting rate of the PCM decreased in the follow­
ing order: pure PCM < nano-PCM < PCM with metal foam < PCM with fins. The same
improving effect (in terms of the liquid fraction) for the case of metal foam was obtained pre­
viously in [49]. Despite the fact that metal foam showed the best cooling performance (as
observed in the previous section), fins provided the best melting rate. At 10000 s, the PCM
melt fractions were 0.25 for pure PCM, 0.27 for nano-PCM, 0.33 for PCM with metal foam,
and 0.41 for PCM with fins. For the surface heat flux and surface heat transfer coefficient
curves, there was a sudden increase during the first 400 s, followed by a rapid decrease,
Figure 4(a and b). The peak maximum for the surface heat flux was 2750 W/m2 for PCM with
metal foam, 1800 W/m2 for nano-PCM and pure PCM, and 1700 W/m2 for PCM with fins.
The peak maximum for the surface heat transfer coefficient was 410 W/m2 K for PCM with
metal foam, 60 W/m2 K for nano-PCM, and 50 W/m2 K for PCM with fins and pure PCM.
Linking these findings with the previous observations on the activated carbon bed, it can be
concluded that the use of enhancers (nanoparticles, metal foam, and fins) in the PCM led to a
higher heat transfer rate between the activated carbon bed and the PCM. This is evidenced by
the higher surface heat flux and surface heat transfer coefficient observed for all the PCM-
enhancer systems compared to pure PCM.
The difference in melting efficiency between metal foam and fins can be explained by their dif­
ferences in surface area and contact with the PCM, as previously mentioned. Metal foam typically
has a higher surface area compared to fins, which allows for more contact with the PCM, result­
ing in better cooling performance. In the study mentioned earlier, the PCM with metal foam
showed the best cooling performance with a peak surface heat flux of 2750 W/m2 and peak sur­
face heat transfer coefficient of 410 W/m2 K. This is higher than the values observed for the
nano-PCM and pure PCM (peak surface heat flux of 1800 W/m2 and peak surface heat transfer
coefficient of 60 W/m2 K) and for the PCM with fins (peak surface heat flux of 1700 W/m2 and
peak surface heat transfer coefficient of 50 W/m2 K). However, despite its better cooling perform­
ance, the efficiency of metal foam in melting the PCM was lower than that of fins. At 10,000 s,
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 15

Figure 5. Velocity contour for the three investigated mass transfer enhancers and the pure paraffin, for different operating times
(400 s, 5000 and 10000 s).

the PCM melt fraction for the PCM with metal foam was 0.33, while the PCM melt fraction for
the PCM with fins was 0.41. This can be attributed to the fact that fins provide better heat trans­
fer from the PCM to the surrounding environment, resulting in a higher melting rate. The higher
16 A. CHIBANI ET AL.

Figure 5. Continued.

heat transfer coefficient of fins facilitates the transfer of heat from the PCM to the surrounding
environment, resulting in a higher rate of melting. Therefore, while metal foam may provide bet­
ter cooling performance, fins are more efficient in melting the PCM. This highlights the impor­
tance of considering multiple factors when selecting enhancers for PCM applications, as the best
performance will depend on the specific application and desired outcome.
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 17

On the other hand, the lower performance of nano-PCM compared to MF and fins enhancers
can be attributed to several factors. Firstly, as observed in the previous section, the thermal con­
ductivity of nanoparticles is generally lower than that of bulk materials. Therefore, the nanopar­
ticles may not be as effective in transferring heat from the activated carbon to the PCM, resulting
in lower melting rates. Secondly, the surface area of nanoparticles is typically much smaller than
that of metal foam or fins. As previously discussed, a larger surface area results in more contact
between the enhancer and the PCM, which can improve the cooling performance of the activated
carbon bed. Therefore, the smaller surface area of nanoparticles could also contribute to the lower
melting rates observed for the nano-PCM. Finally, the size and shape of nanoparticles can also
affect their performance as enhancers. For instance, in a study by Venkateshwar et al. [76], the
impact of various factors, such as size, density, and concentration of nanoparticles, as well as vis­
cosity and the difference in density between the solid and liquid phases of phase change materi­
als, were investigated to determine their effect on the stability of nano-PCMs. The researchers
analyzed two different concentrations of copper oxide (<50 nm) and iron oxide (50–100 nm)
nanoparticles in Rubitherm35 HC and CuO in coconut oil. The experimental results indicated
that nanoparticle size had a more significant effect on stability than nanoparticle density. The sta­
bility of nano-PCMs was improved by the higher density and viscosity of the PCM; however, a
considerable difference in density between the solid and liquid phases of the PCM led to
increased sedimentation. Increasing the concentration of nanoparticles also substantially increased
sedimentation. Therefore, the specific properties of the nanoparticles used in our study may have
also contributed to the lower performance of the nano-PCM compared to MF and fins enhancers.
Overall, the lower melting rate of nano-PCM compared to MF and fins enhancers can be
explained by a combination of lower thermal conductivity, smaller surface area, and possibly
other specific properties of the nanoparticles used in this study.

4.3. Velocity contours


In addition to analyzing the thermal behavior and hydrogen adsorption capacity of the activated
carbon bed in the presence of different PCM-enhancer systems (previous section), we also studied
the velocity contours in the system at different operating times (Figure 5). These contours provide
a visual representation of the flow field around the activated carbon bed and indicate the regions
of high and low velocity. By comparing the velocity contours for the different enhancer systems,
we can gain insights into the hydrodynamic and kinetic behavior of hydrogen charging on acti­
vated carbon in the presence of PCM-enhancer systems, which can help inform the design and
optimization of such systems for practical applications. Furthermore, the velocity contours also
provide insights into the distribution of the enhancer particles in the PCM, which can affect the
overall performance of the system.
Specifically, we examined the velocity contours at 400 s, which corresponds to the time when
the mean hydrogen concentration became constant and the mean activated carbon temperature is
maximal, as well as at 5000s and 10000 s, which represent the middle and later stages of the pro­
cess operation (Figure 5).
The observations of the velocity contours for the pure PCM and all PCM-enhancing systems
indicate that the flow behavior is mainly influenced by the activated carbon bed. At 400 s (1st col­
umn of Figure 5), the high gas velocity at the bottle entrance and its gradual decrease in the lon­
gitudinal direction can be attributed to the resistance of the activated carbon bed to the flow.
This behavior is consistent with previous studies on packed beds, where the velocity profile shows
a similar trend due to the presence of the bed particles [77]. At this time, the velocity vectors are
denser at the top, which is likely due to the buoyancy-driven flow caused by the temperature gra­
dient in the system. The low density at the bottom can be attributed to the higher pressure drop
caused by the resistance of the activated carbon bed. The low-velocity density observed in the
18 A. CHIBANI ET AL.

Figure 6. Mass fraction and temperature contours for the three investigated mass transfer enhancers and the pure paraffin, at
5000 s.

different PCM systems at 400 s can be attributed to the low liquid fraction of the PCM (5%
max.), which is mostly located at the shell wall. This behavior is consistent with previous studies
on solid-liquid phase change systems, where the velocity field is low in the solid region and
higher in the liquid region [70].
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 19

Figure 7. Melting enhancement (%) for the three investigated mass transfer enhancers compared to the pure paraffin, over the
operating period from 500 to 10000 s.

At 5000s (2nd column of Figure 5), the velocity density of the gas in the activated carbon
becomes lighter and more homogenous throughout the bottle. This behavior can be attributed to
the gradual saturation of the activated carbon bed with hydrogen, which leads to a lower resist­
ance to the flow. On the other hand, the relatively higher velocity field observed at the PCM wall
neighboring the AC-bed can be attributed to the presence of the enhancer particles, which can
affect the flow behavior in the PCM. A liquid shell of PCM is formed at this zone, due to the
heat adsorption from the exothermic hydrogen absorption. However, because of the low PCM
conductivity, the progress of the liquid shell of the PCM could be very slow, which is confirmed
by the third column of Figure 3. At 10000 s, the layer thickness of the PCM liquid adjacent to the
activated carbon bed increases a little, leading to an intense velocity field in this region. This
behavior can be attributed to the heat released during the hydrogen adsorption process, which
causes the PCM to melt and increases the liquid fraction. This is consistent with previous studies
on phase change materials, which have shown that the liquid fraction of the PCM can signifi­
cantly affect the velocity field and flow behavior [77]. Based on our observations, it can be con­
cluded that the presence of enhancers in the PCM has a marginal effect on the velocity field in
the PCM side. This could be attributed to the lower thermal conductivity of the PCM compared
to the activated carbon bed, which limits the transfer of thermal energy from the bed to the
PCM. In addition, the presence of enhancers may also contribute to the increase in viscosity of
the PCM (in the case of a nano-PCM system), further hindering the fluid flow and reducing the
velocity field.
In summary, the observations of the velocity contours for the pure and PCM-enhancer systems
indicate that the flow behavior is mainly influenced by the activated carbon bed, with the pres­
ence of enhancer particles affecting the flow behavior in the PCM. The changes in velocity dens­
ity over time can be attributed to the saturation of the activated carbon bed with hydrogen and
the melting of the PCM due to the heat released during the hydrogen adsorption process.

4.4. Mass fraction and temperature contours


For more clarification about the performance of the different hybridized systems (MF-PCM, Fins-
PCM, Nano-PCM) compared to the pure PCM, in Figure 6, the mass fraction and temperature
contours are provided at 5000s, i.e. during the cooling period. These contours provide more
details compared to the velocity contours (Figure 5). As can be seen in the 1st and 2nd columns
of Figure 6 (pure PCM and Nano-PCM systems), a negligible effect is retrieved for the addition
20 A. CHIBANI ET AL.

of nanoparticles (CuO) on the melt fraction of PCM. This is mainly attributed to the low thermal
conductivity of the adopted metal oxide (Table 1). The low efficiency of nanoparticles is clearly
confirmed according to the enhancement plot in Figure 7, where a maximal improvement of
1.2% is retrieved in this case (i.e. metal oxide). Accordingly, a slight enhancement is observed for
the temperature contours with a difference of only �2K (1st and 2nd columns of Figure 6). For
example, at the bed core, the peak temperature in the presence and absence of metal oxide is
372.4 and 375.41 K, respectively. This slight improvement in the heat evacuation is the same over
the whole solid adsorbed; with approximately a constant temperature of 295.15 K throughout the
cooling systems (i.e. pure PCM and Nano-PCM).
On the other side, in the 3rd and 4th columns of Figure 6, remarkable enhancements of the
mass fractions and temperature distributions are observed with the integration of the metal foam
and fins. As it was discussed previously and in corroboration to the findings of Figure 4(a), the
use of fins is more beneficial for a maximal melting of PCM compared to the case of copper
metal foam. Over the majority of the Fins-PCM system, an average of �60% PCM melting is
obtained compared to �40% for metal foam.
This outcome is evidenced through Figure 7, with 15.9% maximal enhancement using fins
compared to 9.1% for metal foam.
In contrast, with the use of metal foam (3rd column in Figure 6), more cooling is obtained
compared to the fins case (4th column in Figure 6). As a result, maximal temperatures of �320
and 329 K are retrieved at the solid bed center for the metal foam and fins system, respectively.
The relatively rapid cooling of the solid adsorbent is translated by more heating of the surround­
ing PCM-metal foam compared to the fins-PCM system (3rd and 4th columns in Figure 6).

5. Conclusions
In conclusion, this study has successfully compared the melting efficiency of phase change materials
with and without enhancers, namely nanoparticles, metal foam, and fins. The results have shown
that the addition of enhancers can significantly improve the melting efficiency of PCM (in the annu­
lar space of the adsorbing bed) in addition to the cooling process of the activated carbon beds.
However, a negligible impact was observed for the adsorbed amount of hydrogen in the presence of
these enhancers. On the other hand, the performance of each enhancer was found to be highly
dependent on their surface area and contact with the PCM. Based on the results, fins have proven to
be the most efficient enhancer for melting PCM, while metal foam has shown better cooling per­
formance. The numerical simulations using ANSYS Fluent have also provided valuable insights into
the velocity field and temperature distribution within the activated carbon bed during hydrogen
charging. Furthermore, the hydrogen adsorption results indicate that activated carbon can be an effi­
cient hydrogen storage material. As a perspective, this study could inspire further research on the
development of novel or the optimization of the existing PCM-enhancer systems for efficient ther­
mal management applications at a large scale. On the other side, another area of research could be
to explore the use of different types of activated carbon (under different operating conditions of
pressure and temperature), such as chemically modified or nanoporous carbon, to enhance hydro­
gen adsorption capacity and improve the overall performance of the system.

Disclosure statement
No potential conflict of interest was reported by the author(s).

ORCID
Atef Chibani http://orcid.org/0000-0002-5861-7498
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 21

References
0[1] I. Dincer and C. Acar, “Review and evaluation of hydrogen production methods for better sustainability,”
Int. J. Hydrogen Energy., vol. 40, no. 34, pp. 11094–11111, 2015. DOI: 10.1016/j.ijhydene.2014.12.035.
0[2] M. Balat, “Possible methods for hydrogen production,” Energy Sources, Part A. Recover. Util. Environ. Eff.,
vol. 31, no. 1, pp. 39–50, Dec. 2008. DOI: 10.1080/15567030701468068.
0[3] R. Kothari, D. Buddhi and R. L. Sawhney, “Comparison of environmental and economic aspects of various
hydrogen production methods,” Renew. Sustain. Energy Rev., vol. 12, no. 2, pp. 553–563, 2008. DOI: 10.
1016/j.rser.2006.07.012.
0[4] E. C. S. Transactions., “(Invited) Novel Fuel Production Based on Sonochemistry and
Sonoelectrochemistry,” Electrochem. Soc., vol. 92, pp. 10, 2019.
0[5] I. K. Kapdan and F. Kargi, “Bio-hydrogen production from waste materials,” Enzyme Microb. Technol, vol.
38, no. 5, pp. 569–582, 2006. Mar. 2018. DOI: 10.1016/j.enzmictec.2005.09.015.
0[6] P. Nikolaidis and A. Poullikkas, “A comparative overview of hydrogen production processes,” Renew.
Sustain. Energy Rev., vol. 67, pp. 597–611, 2017. DOI: 10.1016/j.rser.2016.09.044.
0[7] A. M. Abdalla, S. Hossain, O. B. Nisfindy, A. T. Azad, M. Dawood and A. K. Azad, “Hydrogen production,
storage, transportation and key challenges with applications: a review,” Energy Convers. Manag., vol. 165,
pp. 602–627, 2018. DOI: 10.1016/j.enconman.2018.03.088.
0[8] S. Niaz, T. Manzoor and A. H. Pandith, “Hydrogen storage: materials, methods and perspectives,” Renew.
Sustain. Energy Rev., vol. 50, pp. 457–469, 2015. DOI: 10.1016/j.rser.2015.05.011.
0[9] A. Chibani, “Etude du Stockage et du D�estockage d’Hydrog�ene dans un R�eservoir d’Hydrures �a Triple
Tube Concentrique,” University Batna II., pp. 1–117, 2020.
[10] F. S. Yang, G. X. Wang, Z. X. Zhang, X. Y. Meng and V. Rudolph, “Design of the metal hydride reactors –
A review on the key technical issues,” Int. J. Hydrogen Energy, vol. 35, no. 8, pp. 3832–3840, 2010. DOI: 10.
1016/j.ijhydene.2010.01.053.
[11] A. Chibani, S. Merouani, N. Gherraf and Y. Benguerba, “Thermodynamics and kinetics analysis of hydro­
gen absorption in large-scale metal hydride reactor coupled to phase change material-metal foam-based
latent heat storage system,” Int. J. Hydrogen Energy., vol. 47, no. 64, pp. 27617–27632, 2022. DOI: 10.1016/
j.ijhydene.2022.06.079.
[12] L. Ouyang, et al., “Magnesium-based hydrogen storage compounds: a review,” J. Alloys Compd., vol. 832,
pp. 154865, 2020. DOI: 10.1016/j.jallcom.2020.154865.
[13] K. Møller, et al., “Complex Metal Hydrides for Hydrogen, Thermal and Electrochemical Energy Storage,”
Energies., vol. 10, no. 10, pp. 1645, 2017. DOI: 10.3390/en10101645.
[14] H. Ben M^aad, A. Miled, F. Askri and S. Ben Nasrallah, “Numerical simulation of absorption-desorption
cyclic processes for metal-hydrogen reactor with heat recovery using phase-change material,” Appl. Therm.
Eng., vol. 96, pp. 267–276, 2016. DOI: 10.1016/j.applthermaleng.2015.11.093.
[15] F. Wang, et al., “Entropy optimized flow of Darcy-Forchheimer viscous fluid with cubic autocatalysis chem­
ical reactions,” Int. J. Hydrogen Energy., vol. 47, no. 29, pp. 13911–13920, 2022. DOI: 10.1016/j.ijhydene.
2022.02.141.
[16] S. Mellouli, N. Ben Khedher, F. Askri, A. Jemni and S. Ben Nasrallah, “Numerical analysis of metal hydride
tank with phase change material,” Appl. Therm. Eng., vol. 90, pp. 674–682, 2015. DOI: 10.1016/j.applther­
maleng.2015.07.022.
[17] S. Mellouli, E. Abhilash, F. Askri and S. Ben Nasrallah, “Integration of thermal energy storage unit in a
metal hydride hydrogen storage tank,” Appl. Therm. Eng., vol. 102, pp. 1185–1196, 2016. DOI: 10.1016/j.
applthermaleng.2016.03.116.
[18] S. Mellouli, F. Askri, E. Abhilash and S. Ben Nasrallah, “Impact of using a heat transfer fluid pipe in a
metal hydride-phase change material tank,” Appl. Therm. Eng., vol. 113, pp. 554–565, 2017. DOI: 10.1016/j.
applthermaleng.2016.11.065.
[19] H. El Mghari, 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.
[20] R. Chahine, J. Xiao, R. Peng, D. Cossement and P. Be, “CFD model charge discharge cycle adsorptive
hydrogen storage activated carbon,” Int. J. Hydrogen. Energy., vol. 8, pp. 0–9, 2012. DOI: 10.1016/j.ijhy­
dene.2012.10.119.
[21] M. Hirscher and B. Panella, “Nanostructures with high surface area for hydrogen storage,” J. Alloys
Compd., vol. 404–406, no. SPEC. ISS., pp. 399–401, 2005. DOI: 10.1016/j.jallcom.2004.11.109.
[22] Y. Kojima, et al., “Hydrogen adsorption and desorption by carbon materials,” J. Alloys Compd., vol. 421,
no. 1–2, pp. 204–208, 2006. DOI: 10.1016/j.jallcom.2005.09.077.
22 A. CHIBANI ET AL.

[23] R. Paggiaro, F. Michl, P. B�enard and W. Polifke, “Cryo-adsorptive hydrogen storage on activated carbon.
II: investigation of the thermal effects during filling at cryogenic temperatures,” Int. J. Hydrogen Energy.,
vol. 35, no. 2, pp. 648–659, 2010. DOI: 10.1016/j.ijhydene.2009.11.013.
[24] G. Momen, R. Jafari and K. Hassouni, “On the effect of process temperature on the performance of acti­
vated carbon bed hydrogen storage tank,” Int. J. Therm. Sci., vol. 49, no. 8, pp. 1468–1476, 2010. DOI: 10.
1016/j.ijthermalsci.2010.03.002.
[25] J. Xiao, et al., “Finite element simulation for charge e discharge cycle of cryo-adsorptive hydrogen storage
on activated carbon,” Int. J. Hydrogen Energy., vol. 7, pp. 2–14, 2012. DOI: 10.1016/j.ijhydene.2012.05.079.
[26] G. Hermosilla-Lara, G. Momen, P. H. Marty, B. L. Neindre and K. Hassouni, “Hydrogen storage by adsorp­
tion on activated carbon: investigation of the thermal effects during the charging process,” Int. J. Hydrogen
Energy., vol. 32, no. 10–11, pp. 1542–1553, 2007. DOI: 10.1016/j.ijhydene.2006.10.048.
[27] J. Xiao, L. Tong, C. Deng, P. B�enard and R. Chahine, “Simulation of heat and mass transfer in activated
carbon tank for hydrogen storage,” Int. J. Hydrogen Energy., vol. 35, no. 15, pp. 8106–8116, 2010. DOI: 10.
1016/j.ijhydene.2010.01.021.
[28] E. F. Abbas and S. S. Mohammed, “Study the Effect of Granules Type of The Porous Medium on The Heat
Transfer Enhancement for Double Pipe Heat Exchanger,” Tikrit J. Eng. Sci., vol. 26, no. 4, pp. 43–49, 2019.
DOI: 10.25130/tjes.26.4.07.
[29] A. A. R. Alkumait, T. K. Ibrahim, M. H. Zaidan and A. T. Al, “Thermal and hydraulic characteristics of
TiO 2/water nanofluid flow in tubes possessing internal trapezoidal and triangular rib shapes,” J. Therm.
Anal. Calorim., vol. 147, no. 1, pp. 379–392, 2020. DOI: 10.1007/s10973-020-10289-7.
[30] T. K. Ibrahim, F. Basrawi, M. N. Mohammed and H. Ibrahim, “Effect of perforation area on temperature
distribution of the rectangular fins under natural convection,” ARPN J. Eng. Appl. Sci., vol. 11, no. 10, pp.
6371–6375, 2016.
[31] I. T. K, M. M. Rahman, A. O. M. B. Firdaus and M. Rizalman, “Optimum performance enhancing strat­
egies of the gas turbine based on the effective temperatures,” MATEC Web Conf. Temp., vol. 2, pp. 1–9,
2016. DOI: 10.1051/matecconf2016830002.
[32] S. M. Khalaf, A. A. Mohammed and J. Mohammed, “An experimental investigation on heat transfer
enhancement in an annulus with rotating outer cylinder using nano fluids,” Tikrit j. Eng. Sci., vol. 29, no.
2, pp. 51–60, 2022. DOI: 10.25130/tjes.29.2.7.
[33] T. K. Ibrahim, A. T. Al-Sammarraie, M. S. M. Al-Jethelah, W. H. Al-Doori, M. Reza and H. Tao, “The impact
of square shape perforations on the enhanced heat transfer from fins: experimental and numerical study,” Int. J.
Therm. Sci., vol. 149, no 2019, pp. 106144, Mar. 2020. DOI: 10.1016/j.ijthermalsci.2019.106144.
[34] S. Shafiee and M. Helen, “Different reactor and heat exchanger configurations for metal hydride hydrogen
storage systems a review,” Int. J. Hydrogen Energy., vol. 41, pp. 1–9, 2016. DOI: 10.1016/j.ijhydene.2016.03.
133.
[35] Z. Wu, F. Yang, L. Zhu, P. Feng, Z. Zhang and Y. Wang, “Improvement in hydrogen desorption perform­
ances of magnesium based metal hydride reactor by incorporating helical coil heat exchanger,” Int. J.
Hydrogen Energy., vol. 41, no. 36, pp. 16108–16121, 2016. DOI: 10.1016/j.ijhydene.2016.04.224.
[36] 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.
[37] S. L. Garrison, et al., “Optimization of internal heat exchangers for hydrogen storage tanks utilizing metal
hydrides,” Int. J. Hydrogen Energy., vol. 37, no. 3, pp. 2850–2861, 2012. DOI: 10.1016/j.ijhydene.2011.07.044.
[38] S. Srinivasa Murthy, “Heat and mass transfer in solid state hydrogen storage: a review,” J. Heat Transfer,
vol. 134, no. 3, pp. 031020, 2012. DOI: 10.1115/1.4005156.
[39] S. Izadi, T. Armaghani, R. Ghasemiasl, A. J. Chamkha and M. Molana, “A comprehensive review on mixed
convection of nanofluids in various shapes of enclosures,” Powder Technol, vol. 343, pp. 880–907, 2019.
DOI: 10.1016/j.powtec.2018.11.006.
[40] K. M. Khanafer and A. J. Chamkha, “Hydromagnetic natural convection from an inclined porous square
enclosure with heat generation,” Numer. Heat Transf. Part A Appl, vol. 33, no. 8, pp. 891–910, Jun. 1998.
DOI: 10.1080/10407789808913972.
[41] M. Ghalambaz, A. Doostani, E. Izadpanahi and A. J. Chamkha, “Conjugate natural convection flow of Ag–
MgO/water hybrid nanofluid in a square cavity,” J. Therm. Anal. Calorim., vol. 139, no. 3, pp. 2321–2336,
2020. DOI: 10.1007/s10973-019-08617-7.
[42] M. Ghalambaz, M. Aljaghtham, A. J. Chamkha, M. Fteiti and A. Abdullah, “Latent heat thermal energy
storage in a shell-tube: a wavy partial layer of metal foam over tubes,” J. Energy Storage., vol. 59, pp.
106493, 2023. DOI: 10.1016/j.est.2022.106493.
[43] H. M. Sadeghi, M. Babayan and A. Chamkha, “Investigation of using multi-layer PCMs in the tubular heat
exchanger with periodic heat transfer boundary condition,” Int. J. Heat Mass Transf., vol. 147, pp. 118970,
2020. DOI: 10.1016/j.ijheatmasstransfer.2019.118970.
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 23

[44] M. Ghalambaz, M. Aljaghtham, A. J. Chamkha, A. Abdullah, I. Mansir and M. Ghalambaz, “Mathematical


modeling of heterogeneous metal foams for phase-change heat transfer enhancement of latent heat thermal
energy storage units,” Appl. Math. Model., vol. 115, pp. 398–413, 2023. DOI: 10.1016/j.apm.2022.10.018.
[45] M. Ghalambaz, M. Aljaghtham, A. J. Chamkha, A. Abdullah, U. Alqsair and M. Ghalambaz, “Dynamic
melting in an open enclosure supported by a partial layer of metal foam: a fast thermal charging approach,”
Int. J. Heat Mass Transf., vol. 203, pp. 123760, 2023. DOI: 10.1016/j.ijheatmasstransfer.2022.123760.
[46] V. Joshi and M. K. Rathod, “Thermal transport augmentation in latent heat thermal energy storage system
by partially fi lled metal foam : A novel con fi guration,” J. Energy Storage, vol. 22, pp. 270–282, 2019.
DOI: 10.1016/j.est.2019.02.019.
[47] Y. Cui, J. Xie, J. Liu and S. Pan, “Review of phase change materials integrated in building walls for energy
saving,” Procedia Eng, vol. 121, pp. 763–770, 2015. DOI: 10.1016/j.proeng.2015.09.027.
[48] A. Chibani, G. Mecheri, A. Dehane, S. Merouani and I. Ferhoune, “Performance improvement of adsorptive
hydrogen storage on activated carbon: effects of phase change material and inconstant mass flow rate,” J.
Energy Storage, vol. 56, no pb, pp. 105930, 2022. DOI: 10.1016/j.est.2022.105930.
[49] A. Chibani, G. Mecheri, S. Merouani and A. Dehane, “Hydrogen charging in AX21 activated carbon-PCM-
metal foam-based industrial-scale reactor: numerical analysis,” Int. J. Hydrogen Energy, vol. 48, no. 82, pp.
32025–32038, 2023. DOI: 10.1016/j.ijhydene.2023.03.049.
[50] A. Valan Arasu, A. P. Sasmito and A. S. Mujumdar, “Thermal performance enhancement of paraffin wax
with Al2O3 and CuO nanoparticles - a numerical study,” Front. Heat Mass Transf., vol. 2, no. 4, pp. 1–7,
2011. DOI: 10.5098/hmt.v2.4.3005.
[51] H. Senobar, M. Aramesh and B. Shabani, “Nanoparticles and metal foams for heat transfer enhancement of
phase change materials: a comparative experimental study,” J. Energy Storage., vol. 32, no. May, pp. 101911,
2020. DOI: 10.1016/j.est.2020.101911.
[52] H. Jin, L. Fan, M. Liu, Z. Zhu and Z. Yu, “A pore-scale visualized study of melting heat transfer of a paraf­
fin wax saturated in a copper foam: effects of the pore size,” Int. J. Heat Mass Transf, vol. 112, pp. 39–44,
2017. DOI: 10.1016/j.ijheatmasstransfer.2017.04.114.
[53] L. Asip, M. Mahabat, H. Farooq, M. Irfan, D. Brabazon and I. Ul, “Dominant roles of eccentricity, fin
design, and nanoparticles in performance enhancement of latent thermal energy storage unit,” J. Energy
Storage, vol. 43, no. August, pp. 103181, 2021. DOI: 10.1016/j.est.2021.103181.
[54] 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. Manag, vol. 166,
pp. 241–259, Mar. 2018. DOI: 10.1016/j.enconman.2018.04.016.
[55] 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.
[56] 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 2021, pp. 104449, September 2022. DOI: 10.1016/j.est.
2022.104449.
[57] A. Chibani, S. Merouani, C. Bougriou and L. Hamadi, “International Journal of Heat and Mass Transfer
Heat and mass transfer during the storage of hydrogen in LaNi 5 -based metal hydride : 2D simulation
results for a large scale, multi-pipes fixed- bed reactor,” Int. J. Heat Mass Transf., vol. 147, no. xxxx, pp.
118939, 2019. DOI: 10.1016/j.ijheatmasstransfer.2019.118939.
[58] 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.
[59] 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, pp. 118622, 2022. DOI: 10.1016/j.applthermaleng.2022.118622.
[60] 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.
[61] 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.
[62] F. Ye, J. Xiao, B. Hu, P. Benard and R. Cgahine, “Implementation for model of adsorptive hydrogen storage
using UDF in fluent,” Phys. Procedia., vol. 24, pp. 793–800, 2012. DOI: 10.1016/j.phpro.2012.02.118.
[63] A. Chibani, C. Bougriou and S. Merouani, “Simulation of hydrogen absorption/desorption on metal hydride
LaNi5-H2: mass and heat transfer,” Appl. Therm. Eng., vol. 142, pp. 110–117, 2018. DOI: 10.1016/j.applther­
maleng.2018.06.078.
24 A. CHIBANI ET AL.

[64] J. Xiao, M. Hu, P. B�enard and R. Chahine, “Simulation of hydrogen storage tank packed with metal-organic
framework,” Int. J. Hydrogen Energy., vol. 38, no. 29, pp. 13000–13010, 2013. DOI: 10.1016/j.ijhydene.2013.
03.140.
[65] H. Lani, A. Chibani and C. Bougriou, “Effect of the tank geometry on the storage and destocking of hydro­
gen on metal hydride,” Int. J. Hydrogen Energy,., vol. 42, no. 36, pp. 23035–23044, 2017. DOI: 10.1016/j.
ijhydene.2017.07.102.
[66] M. Afzal, N. Gupta, A. Mallik, K. S. Vishnulal and P. Sharma, “Experimental analysis of a metal hydride
hydrogen storage system with hexagonal honeycomb-based heat transfer enhancements-part B,” Int. J.
Hydrogen Energy., vol. 46, no. 24, pp. 13131–13141, 2021. DOI: 10.1016/j.ijhydene.2020.11.275.
[67] J. Xiao and B. Pierre, “Charge-discharge cycle thermodynamics for compression hydrogen storage system,”
Hydrog. Energy., vol. 41, pp. 1–9, 2016. DOI: 10.1016/j.ijhydene.2015.12.136.
[68] J. Xiao, R. Peng, D. Cossement, P. B�enard and R. Chahine, “Heat and mass transfer and fluid flow in cryo-
adsorptive hydrogen storage system,” Int. J. Hydrogen Energy., vol. 38, no. 25, pp. 10871–10879, 2013. DOI:
10.1016/j.ijhydene.2013.04.042.
[69] R. Chahine, J. Xiao, L. Tong, D. Cossement and P. Be, “cfd simulation charge discharge cycle cryo-adsorp­
tive hydrogen storage on activated carbon, ” vol. 7, pp. 12893–12904, 2012. DOI: 10.1016/j.ijhydene.2012.
05.079.
[70] A. Chibani, S. Merouani, N. Gherraf, I. Ferhoune and Y. Benguerba, “Numerical investigation of heat and
mass transfer during hydrogen desorption in a large-scale metal hydride reactor coupled to a phase change
material with nano-oxide additives,” Int. J. Hydrogen Energy, vol. 47, no. 32, pp. 14611–14627, 2022. DOI:
10.1016/j.ijhydene.2022.02.171.
[71] 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 Transf., vol. 144, pp. 118619, 2019. DOI: 10.1016/j.ijheat­
masstransfer.2019.118619.
[72] J. M. Mahdi and E. C. Nsofor, “Multiple-segment metal foam application in the shell-and-tube PCM ther­
mal energy storage system,” J. Energy Storage., vol. 20, pp. 529–541, June 2018. DOI: 10.1016/j.est.2018.09.
021.
[73] Z. A. Qureshi, E. Elnajjar, O. Al-Ketan, R. A. Al-Rub and S. B. Al-Omari, “Heat transfer performance of a
finned metal foam-phase change material (FMF-PCM) system incorporating triply periodic minimal surfa­
ces (TPMS),” Int. J. Heat Mass Transf, vol. 170, pp. 121001, 2021. DOI: 10.1016/j.ijheatmasstransfer.2021.
121001.
[74] 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. 2019, pp. 113993, 2020. DOI: 10.1016/j.apenergy.2019.113993.
[75] A. Chibani, S. Merouani, C. Bougriou and L. Hamadi, “Heat and mass transfer during the storage of hydro­
gen in LaNi5-based metal hydride: 2D simulation results for a large scale, multi-pipes fixed-bed reactor,”
Int. J. Heat Mass Transf., vol. 147, pp. 118939, Feb. 2020. DOI: 10.1016/j.ijheatmasstransfer.2019.118939.
[76] K. Venkateshwar, N. Joshy, H. Simha and S. Mahmud, “Quantifying the nanoparticles concentration in
nano-PCM,” J Nanopart Res., vol. 21, no. 12, pp. 260, 2019. DOI: 10.1007/s11051-019-4716-x.
[77] S. Kyoung, S. Ferekh, G. Gwak, A. Jo and H. Ju, “Three-dimensional modeling and simulation of hydrogen
desorption in metal hydride hydrogen storage vessels,” Int. J. Hydrogen Energy., vol. 40, no. 41, pp. 14322–
14330, 2015. DOI: 10.1016/j.ijhydene.2015.03.114.

You might also like