1 s2.0 S1359431120329744 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

Applied Thermal Engineering 178 (2020) 115492

Contents lists available at ScienceDirect

Applied Thermal Engineering


journal homepage: www.elsevier.com/locate/apthermeng

Nano-enhanced phase change materials (NePCMs): A review of numerical T


simulations

Teng Xiong, Long Zheng, Kwok Wei Shah
Department of Building, School of Design and Environment, National University of Singapore, Singapore 117566, Singapore

H I GH L IG H T S

• Numerical methods and prediction models are summarized and discussed.


• Melting and solidification of NePCMs inside different containers are reviewed.
• The effects of nanostructure type, morphology, size, and concentration are analyzed.
• The reasons for the discrepancy between simulation and experiment are identified.
• There is a trade-off between conduction enhancement, convection degradation, and latent heat reduction.

A R T I C LE I N FO A B S T R A C T

Keywords: Dispersing thermally conductive nanostructures is an effective method to improve the thermal performance of
Nano-enhanced phase change materials phase change materials (PCMs). For this purpose, nanocarbons, nanometals, and nano metal oxides have been
Numerical methods used to develop nano-enhanced phase change materials (NePCMs) with unique thermal properties. However,
Prediction models review papers focusing on the numerical simulations of NePCMs are still scarce. The present review provides a
Nanostructures
comprehensive overview of the latest numerical studies on NePCMs for thermal energy storage (TES). These
Melting and solidification
studies are mainly based on single-phase approaches, and the simulation results largely depend on the used
Thermal energy storage
prediction models of effective thermophysical properties. Accordingly, the most common numerical methods
and prediction models are reviewed to address their advantages and limitations. Then, the focus is placed on
melting and solidification of NePCMs inside different containers, including rectangular cavities, tubes, cylinders,
spheres, and annulus. In-depth insights are given into the effects of nanostructure type, morphology, size, and
concentration on heat storage and release performance. The pros and cons of dispersing nanoparticles and other
heat transfer enhancement techniques are also compared, such as mounting fins and using porous foams.
Moreover, a critical discussion is presented to identify the reasons for the discrepancy between simulation and
experiment, as well as the research gaps and future directions. This review aims to update the existing NePCM
studies using different simulation techniques, and to reveal the basic phase change behavior of NePCMs from the
reported results.

T temperature (K) S momentum source term


Ceff effective heat capacity (J/kg·K) Amush mushy zone constant
k thermal conductivity (W/m·K) ui velocity (m/s)
Cp heat capacity (J/kg·K) A intrinsic viscosity (kg/m·s)
L latent heat of fusion (J/kg) d diameter (m)
H total enthalpy (J/kg) M molecular weight (kg/kmol)
m mass (kg) N Avogadro number (mol-1)
t time (s) C’ empirical constant
l length (m) B Boltzmann constant (J/K)
f liquid fraction γ fraction of liquid volume


Corresponding author.
E-mail address: bdgskw@nus.edu.sg (K.W. Shah).

https://doi.org/10.1016/j.applthermaleng.2020.115492
Received 6 February 2020; Received in revised form 2 May 2020; Accepted 14 May 2020
Available online 26 May 2020
1359-4311/ © 2020 Elsevier Ltd. All rights reserved.
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

Re Reynolds number SDBS sodium dodecylbenzene sulfonate


Pr Prandtl number SDS sodium dodecyl sulfate
Lii geometrical factor UV-vis Ultraviolet-visible spectrophotometry
a aspect ratio PIV particle image velocimetry
kii equivalent thermal conductivity (W/m·K) DSC differential scanning calorimetry
R thermal boundary resistance (m2·K/W)
Gr Grashof number 1. Introduction
Ra Rayleigh number
Ha Hartmann number Energy poverty is a major challenge facing the world today due to
Nu Nusselt number the fact that about 850 million people still lack access to electricity.
Gr Grashof number Global energy demand is projected to increase by roughly 30% between
St Stefan number today and 2040, with growing demand for energy services and effi-
subscript ciency improvements [1]. Countries with fast-growing economies,
especially Asian countries, are largely responsible for the growth [2]. So
s solid far, the share of fossil fuels (i.e. oil, coal, and natural gas) in the world’s
l liquid primary energy sources has remained stubbornly high at 85%. Without
0 initial a doubt, the burning of fossil fuels will greatly contribute to greenhouse
np NePCM gas emissions, leading to global warming, extreme weather events, sea
p PCM level rise, and other long-term impacts. According to a new report from
n nanoparticle the International Energy Agency [3], the energy- and process-related
e equivalent CO2 emissions hit another record high of 33 gigatons in 2019, following
ref reference two years of increases. The world must therefore balance the role of
Greek letters energy in social and economic development with the need to dec-
arbonize, and transition from a fossil fuel dominated energy mix to a
ρ density (kg/m3) low-carbon one. For this purpose, many research works have been
ε a small number conducted to tap into secure, affordable and renewable thermal energy
µ dynamic viscosity (kg/m·s) sources (e.g. solar thermal, geothermal, and ocean thermal). The
ϕ volume fraction (vol.%) drawback of the intermittent nature of renewables can be overcome by
ϕmax maximum packing factor installing a thermal energy storage (TES) system between the energy
β expansion coefficient (1/K) supply and demand sectors.
δ thickness (m) TES can be based on sensible heat, latent heat, and thermochemical
List of abbreviations heat. Among these methods, latent heat storage using phase change
materials (PCMs) has received the most attention in the past decade.
PCMs phase change materials PCMs are substances that absorb and release a lot of thermal energy by
NePCMs nano-enhanced phase change materials melting and freezing. The advantages of PCMs come from various as-
TES thermal energy storage pects, including high TES density, narrow operating temperature span,
PV photovoltaic and a small volume change during the phase transition [4,5]. In some
CSP concentrated solar power climate regions, PCMs were found to be effective in improving the
FVM finite volume method thermal performance and energy efficiency of buildings. PCM-in-
FEM finite element method tegrated wallboards, bricks, ceilings, and roofs with high thermal in-
EHCM effective heat capacity method ertia are able to attenuate the room temperature fluctuation, reduce the
EM enthalpy method cooling and heating energy demand, and downsize the HVAC system in
HSTM heat source term method buildings [6]. These are passive cooling and heating technologies. Si-
CFD computational fluid dynamics milarly, PCMs have also been extensively used for thermal management
LBM lattice-Boltzmann method of photovoltaic (PV) cells [7], Li-ion batteries [8], electronic devices
MD molecular dynamic [9], and vehicle components [10]. An actively charged/discharged
DEMT differential effective medium theory PCM system can serve as a buffer to balance the time and spatial mis-
EMT effective medium theory match between renewable energy sources and consumers. For example,
PG propylene glycol installing storage tanks packed with molten salt PCMs (phase change
EG ethylene glycol temperature range of 300–500 °C) has turned out to be a promising TES
MWCNT multi-walled carbon nanotubes method in commercial concentrated solar power (CSP) plants which
GnP graphene nanoplatelets involve interrupted energy conversion cycles [11]. Another idea is to
CNF carbon nanofibers use PCMs with different melting points for waste heat recovery in
SWCNT single-walled carbon nanotubes thermal power plants [12]. The captured waste heat can be reused by
CNT carbon nanotubes district energy systems to provide a cleaner way for heating buildings.
HTF heat transfer fluid Recent advances in PCMs and their applications in different scenarios
ND nanodiamond can be found in the review articles [13–16].
EPM enthalpy porosity method From a material point of view, the main defect of pure PCMs is their
MADM multi-attribute decision making poor thermal conductivity (usually between 0.2 and 0.7 W/m·K), which
DDF double distribution function limits the heat transfer efficiency of TES systems. Enhancing the
VOF volume of fluid thermal performance of PCMs has therefore become critical to their
LTE local thermal equilibrium effectiveness. With the fast development of nanotechnology and con-
LTNE local thermal non-equilibrium trollable synthesis, an easy-to-use method is to disperse thermally
RSM response surface method conductive nanostructures that are smaller than 100 nm in at least one
SGFEM standard Galerkin finite element method dimension. The commonly used ultra-small nanomaterials are carbons
UDF user defined function (e.g. carbon, graphene, graphite), metals (e.g. Cu, Ag), and metal oxides

2
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

Table 1
Summary of experimental studies on NePCMs.
Nanostructure Morphology Average size Concentration Base PCM Phase change Thermal enhancement Ref.
temperature

graphene nanoplatelet d of 10 μm, δ of 64 nm 5.0 wt% paraffin wax 51–58 °C 164% thermal conductivity ↑ [19]
graphene nanoplatelet d of 5–10 μm, δ of 4–20 nm 10.0 wt% n-eicosane 35.7 °C 400% thermal conductivity ↑ [20]
graphite nanoplatelet d of 15 μm, δ of 10 nm 7.0 wt% paraffin wax 51–55 °C 207% thermal conductivity ↑ [21]
graphite nanoplatelet d of 15 μm, δ of 35 nm 10.0 wt% paraffin wax 29.7 °C 167% thermal conductivity ↑ [22]
carbon nanotube d of 6–9 nm, l of 5 μm 1.0 wt% n-eicosane 32–40 °C 113% thermal conductivity ↑ [23]
copper nanoparticle d of 10–30 nm 0.5 wt% sodium acetate trihydrate 58–62 °C 25% thermal conductivity ↑ [24]
20% heat transfer rate ↑
copper nanoparticle d of 25 nm 1.0 wt% paraffin wax 55–60 °C 30% melting rate ↑ [25]
28% solidification rate ↑
copper nanowire d of 80–150 nm, l of 10 μm 11.9 vol% tetradecanol 38–40 °C 793% thermal conductivity ↑ [26]
copper nanowire d of 50 nm, l of 10 μm 0.17 wt% calcium chloride 28–30 °C 52% thermal conductivity ↑ [27]
hexahydrate
silver nanoparticle d of 10–18 nm 5.0 wt% organic ester 5.8–7.7 °C 67% thermal conductivity ↑ [28]
silver nanowire d of 50–100 nm, l of 19.3 wt% polyethylene glycol 43–60 °C 1130% thermal conductivity [29]
5–20 μm ↑
copper oxide nanoparticle d of 37–59 nm 0.1 wt% water −0.8–0 °C 35% solidification rate ↑ [30]
alumina nanoparticle d of 10–20 nm 10.0 wt% paraffin wax 61.6 °C 43% thermal conductivity ↑ [31]
27% melting rate ↑
titanium dioxide nanoparticle d of 21 nm 5.0 wt% palmitic acid 60–62 °C 80% thermal conductivity ↑ [32]
titanium dioxide nanoparticle d of 20 nm 1.13 vol% barium chloride hydrate −9.3 °C 13% thermal conductivity ↑ [33]
titanium dioxide nanoparticle d of 50 nm 0.3 wt% paraffin wax 55–58 °C 25% thermal conductivity ↑ [34]
copper oxide nanoparticle d of 40 nm 55–59 °C 29% thermal conductivity ↑
graphene oxide nanosheet l of 5–20 μm, δ of 50 nm 56–57 °C 101% thermal conductivity ↑
magnetite nanoparticle d of 40–75 nm 20.0 wt% paraffin wax 43–50 °C 60% thermal conductivity ↑ [35]
8% latent heat ↑
boron nitride nanosheet l of 0.5–2 μm, δ of 100 nm 10.0 wt% paraffin wax 50–58 °C 60% thermal conductivity ↑ [36]

Note: d, δ, l, and ↑ stand for diameter, thickness, length, and enhancement, respectively.

(e.g. CuO, Al2O3, MgO, TiO2, ZnO) [17]. These highly conductive na- performance of NePCM systems, and one may overestimate the real
nomaterials have different morphologies/shapes, including 0D (e.g. performance using the same indicators for pure PCMs. Hu et al. [45],
particles, diamonds), 1D (e.g. tubes, rods, fibers, wires), and 2D (e.g. for example, emphasized that a faster melting rate does not always
sheets, platelets, flakes). The colloidal mixtures of nanostructures and represent a higher heat storage rate in the cases with nano dispersion. A
PCMs are called nano-enhanced phase change material (NePCMs), deeper insight should be given into the evaluation methods used in the
which can be used in a number of thermal energy-related applications. NePCM literature.
Table 1 summarizes several noteworthy studies which have successfully To the best of our knowledge, the previous review papers [46–48]
prepared NePCMs with remarkably enhanced thermal conductivities. have not sufficiently discussed the issues mentioned above. Hence, the
The main factors for the thermal enhancement include the type of base novelty of this review is to (1) update the existing NePCM studies using
PCM, as well as the type, morphology, size, and concentration of na- different numerical techniques, (2) compare the assumptions adopted
nostructures [18]. Another important factor is the dispersion stability of in the reported models, (3) address the advantages and limitations of
nanostructures in the base PCM, which is discussed in Section 5.1. each model, (4) identify the reasons for the discrepancy between si-
Besides the thermal conductivity, nano dispersion could dramati- mulation and experiment, and (5) reveal the basic phase change be-
cally change the other thermophysical properties of the base PCM, such havior of NePCMs. Fig. 1 shows an overview of this work. It starts with
as viscosity, density, latent heat of fusion, specific heat, phase change a brief review of the physical models of NePCMs (Section 2.1). Next, the
temperature, and supercooling degree [37]. Some of these changes are most common numerical methods that used in macroscale, mesoscale,
negative and should not be neglected [38]. Numerically simulate the and molecular-scale simulations are introduced (Section 2.2). This is
effects of the changed properties can greatly facilitate the selection of followed by a summary of the prediction models of effective thermo-
suitable nanostructures. The simulation results also provide a cost-ef- physical properties that used in NePCM studies, along with the factors
fective way to better understand the flow and heat transfer character- considered in each model (Section 2.3). Then, we review melting
istics of NePCMs in containers [39]. However, there are also challenges (Section 3) and solidification (Section 4) processes of NePCMs inside
involved. The first difficulty lies in how to properly consider the pre- different containers. The nanostructures are divided into nanocarbons,
sence of nanostructures in numerical parameters and models. Hence, nanometals, and nano metal oxides. Particular attention has been given
there is a necessity to elaborate on the difference between the numer- to the improved thermal performance of NePCM systems. The key
ical simulation of phase change and that with nano dispersion. The findings are summarized in tables for comparison at the end of each
numerical methods for modeling of NePCMs are scattered across the section. In Section 5, we give a critical discussion on the deviations
literature, which requires a detailed review to address the pros and cons between numerical and experimental studies, the basic melting and
of each approach. Secondly, the numerical models of NePCMs should be solidification characteristics of NePCMs, as well as the pros and cons of
validated by experiments. However, some experimental studies of dispersing nanostructures. Based on the literature review, we identify
NePCMs [40–44] have reported results that contradicted other simu- the research gaps in the past studies, and propose some possible di-
lation results, particularly in the melting process. Hence, it is para- rections for future work.
mount to identify the reasons for these discrepancies, and to use ap-
propriate methods to improve the accuracy of simulation results. Since 2. Numerical simulations of NePCMs
numerical models are based on some assumptions to make the calcu-
lation process easier, the existence of nanostructures may require ad- 2.1. Physical models of NePCMs
ditional considerations in the classical phase change formulations.
Furthermore, there is currently no standard way to assess the thermal When a NePCM is fully melted, it can be seen as a nanofluid. The

3
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

Fig. 1. Overview of the present review.

difference is that the liquid NePCM is usually under buoyancy-driven simulations of nanofluids using two-phase mixture models. The nu-
laminar flow condition, while the nanofluid can be under both laminar merical simulations of NePCMs based on two-phase approaches, to the
and turbulent flow conditions [49]. Further, the NePCM also performs best our knowledge, have not yet been reported in the literature. One
in solid and mushy (semi-solid) states, so it will show different thermal reason could be that it is much more challenging to couple the physical
and rheological behavior as its temperature goes up or down. To con- states of NePCMs (i.e. solid, mushy, and liquid) with the dynamics of
sider the dynamics of nanoparticles in the base fluid, the physical nanoparticles (i.e. Brownian motion and thermophoresis) in the two-
models of liquid NePCMs (nanofluids) can be based on single-phase and phase nanofluid models. This means some techniques are needed to
two-phase approaches. track and control the nanoparticle diffusion in the base PCM upon
There are two main single-phase models, namely the homogeneous phase change. For instance, Brownian motion should be ignored in the
and thermal dispersion models. The homogeneous model assumes the solid NePCM, while it should be individually considered in the mushy
ultra-fine nanoparticles are uniformly dispersed in the base fluid. The zone and in the liquid NePCM. Further, the reliability of these techni-
nanoparticles can be easily fluidized, so there is no slip between the ques should be validated against experimental data, which remains
base fluid and nanoparticles [50]. Further, the base fluid and nano- fairly incomplete. In fact, the common two-phase models such as mix-
particles are in thermal equilibrium [51]. Under these assumptions, the ture model and Eulerian model are unable to be used in conjunction
nanofluid can be treated as a homogeneous mixture with average with the solidification and melting model in ANSYS Fluent [62]. Con-
thermophysical properties of components. The homogeneous model is a sidering these reasons, the vast majority of simulations have been based
“static” model, and it has the benefits of simplicity and high compu- on single-phase approaches and classical phase change formulations,
tational efficiency since all governing equations are given in the general along with the following assumptions [63,64]:
form. However, the simulation results will largely depend on the ef-
fective thermophysical properties of nanofluid, particularly the thermal (1) The NePCM is homogeneous and isotropic.
conductivity and viscosity. This is because the influence of dispersing (2) The liquid NePCM is a Newtonian and incompressible fluid.
nanoparticles is only captured in the effective properties. To improve (3) The liquid NePCM is under laminar flow, and the viscous dissipa-
the homogeneous model, several thermal dispersion models [52–54] tion is negligible.
were developed by considering the contribution of thermal dispersion (4) The phase change process is conduction and convection controlled,
to heat transfer enhancement of nanofluid. The thermal dispersion ef- and the Boussinesq approximation is used to account for the natural
fect results from the random and chaotic motion of nanoparticles in the convection.
base fluid, which produces small perturbations in both velocity and (5) The thermophysical properties of NePCM are temperature-in-
temperature [55]. Hence, the thermal dispersion model is a “dynamic” dependent, except the thermal conductivity and the density in the
model. For a fully developed nanofluid flow, it can produce a better buoyance term.
result by further considering the nanoparticle migration, which is
mainly due to Brownian motion and thermophoresis [56]. 2.2. Numerical methods used in simulations of NePCMs
The slip velocity between the base fluid and nanoparticles may not
be zero in the cases with high particle concentrations [57]. Further, the Depending on the scale of the problem, the simulations of NePCMs
nanoparticle migration can lead to nonuniform concentration dis- can be divided into three main types: macroscale, mesoscale, and mo-
tributions and thermal properties throughout the flow domain [58]. lecular-scale. As shown in Fig. 2, each simulation is based on numeri-
The above are the solid-liquid flow features of nanofluids. Hence, in the cally solving a number of mathematical equations for the concerned
two-phase methods, the base fluid and nanoparticles are simulated as physical phenomena. In the literature, the first two simulations have
two individual phases with different velocities and possible different been widely adopted to investigate the flow and heat transfer char-
temperatures [59]. The most common two-phase mixture models are acteristics of NePCMs, while the third one has been used to predict the
Eulerian-Eulerian and Eulerian-Lagrangian. The former is suitable for a effective thermophysical properties of NePCMs. This section introduces
high particle volume fraction, while the latter is suitable for a low the numerical methods used in each simulation.
particle volume fraction [60]. The advantages and limitations of each
model have been well explained in the literature [59,61]. Though two-
2.2.1. Macroscale simulation
phase models may provide a realistic result by considering the slip
Macroscale simulations, such as finite volume method (FVM) and
mechanisms between nanoparticles and base fluid, they are computa-
finite element method (FEM), are based on directly discretizing the
tionally demanding and time consuming.
governing equations for fluid flow and heat transfer. The discrete
Despite numerous studies have focused on the numerical
equations are then numerically solved to obtain velocity and

4
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

Fig. 2. Numerical simulations of NePCMs.

Fig. 3. Flow chart of the macroscale simulation of NePCMs.

temperature fields. Fig. 3 shows a flow chart of the macroscale simu- NePCMs. The correlations can be obtained by prediction models or
lation of NePCMs. Under the assumptions mentioned in Section 2.1, the experimental measurements. As for the phase change (also referred to
basic fundamental is to introduce suitable correlations to capture the as latent heat) models, most simulation studies have used fixed grid
influence of dispersing nanostructures on the material properties of methods which can be divided into three main types: effective heat

5
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

capacity method (EHCM), enthalpy method (EM), and heat source term In this method, Cp and L are combined into the total enthalpy (H) as
method (HSTM). All these methods are straightforward for coding into the primary dependent variable in the energy equation Eq. (4):
FVM or FEM based computational fluid dynamics (CFD) packages (e.g.
∂H
COMSOL Multiphysics and ANSYS Fluent). ρ = ∇ ·(k∇T )
∂t (4)
(1) Effective heat capacity method
In this method, the temperature (T) is the primary dependent Similar to the EHCM, the key of this model is to use a suitable H (T)
variable in the energy equation Eq. (1). relationship, from which the temperature can be derived. The most
common H (T) relationship can be written as [67,73,74]:
∂T
ρCeff = ∇ ·(k∇T )
∂t (1) T
⎧ ∫T0 Cp, s dT T < Ts (solid state)
where k and ρ are the thermal conductivity and density of NePCM, ⎪ T
s T − Ts
H = ∫T Cp, s dT + L T − T Ts ⩽ T < Tl (mushy state)
respectively. Ceff is the effective heat capacity of NePCM which re- ⎨ 0 l s

presents both sensible heat (Cp) and latent heat (L). ⎪ ∫Ts C dT + L + ∫T C dT T ⩾ Tl (liquid state)
p, s p, l
⎩ T0 Tl (5)
The EHCM is easy to implement since it only has one unknown
variable T. The energy equation given in the general form also makes where T0 is the initial temperature at which the H is 0 kJ/kg.
the EHCM suitable for implicit discretization in time [65]. From Eq. (1), It should be noted that the above relationship is a simplified one
it is clear that the key of this method lies in the specification of Ceff (T) assuming the growth of total enthalpy in each state is linear. The real H
relationships which can be mainly categorized into two types: empirical (T) curves should be determined by DSC. As mentioned earlier, the
and experimental relationships. small mass of sample may influence the DSC results. To characterize the
The empirical relationship considers the L as a constant in the Ceff thermal properties of a bulk material, T-history method was proposed
over the phase change temperature range, within which the NePCM is by Zhang et al. [75]. Owing to its simplicity and accuracy, the T-history
treated as a pseudo mushy zone. Thus, the Ceff can be given by [66]: method has been used for thermal characterizations of NePCMs, in-
T < Ts (solid state) cluding specific heat, latent heat of fusion, and thermal conductivity
⎧Cp, s [76–78]. The enthalpy method has gained popularity since it can solve
⎪ Cp, s + Cp, l L
Ceff = + Ts ⩽ T < Tl (mushy state) both smooth and sharp phase transition problems. However, the tem-
⎨ 2 Tl − Ts
⎪Cp, l T ⩾ Tl (liquid state) perature at a typical grid point may oscillate with time [79]. Further,
⎩ (2)
the use of linear H (T) curves could produce discrepancies between
where Cp,s and Cp,l are the specific heat capacities of solid NePCM simulation and experimental results [80].
and liquid NePCM, respectively. (3) Heat source term method
When dealing with narrow or isothermal phase change problems, In fact, the HSTM is an alternative discretization for the enthalpy
the use of the above relationship may create singularity and con- method. By splitting the H into Cp and L, Eq. (4) can be written as [81]:
vergence issues due to the sharp jump of Ceff in the mushy zone. Hence,
∂T ∂f
Jin et al. [67] pointed out the accuracy of using the empirical re- ρCp = ∇ ·(k∇T ) − ρL
∂t ∂t (6)
lationship is acceptable when the phase change temperature range is
greater than 2 °C. Further, a small time step is required for accurate where f is the liquid fraction of NePCM, and it can be written as:
results. This is because when the temperature variation in one time step
is greater than the phase change temperature range, the solver may ⎧0 T < Ts (solid state)
⎪ T − Ts
underestimate the latent heat contribution in the mushy zone [66]. f= Ts ⩽ T < Tl (mushy state)
⎨ Tl − Ts
Hence, this method is less computationally efficient. Some similar Ceff ⎪1 T ⩾ Tl (liquid state)
⎩ (7)
(T) relationships were compared in a recent study [68].
The experimental relationship can be obtained by differential From Eq. (6), the primary dependent variable in the HSTEM is T.
scanning calorimetry (DSC), which is more straightforward and accu- The HSTM can deal with smooth and sharp phase change problems,
rate than the empirical relationship. Based on the DSC result, the Ceff similar to the enthalpy method. The use of the above f (T) relationship
can be given by [69,70]: was found to be effective in achieving a faster convergence rate under a
Newton-Raphson procedure [82]. The key of this method is how the f is
dH ⎛m · dT ⎞
Ceff = updated in each iteration, which usually requires an optimum under-
dt ⎝ dt ⎠ (3)
relaxation factor [66]. A modified HSTEM that does not need any
where H is the total enthalpy of NePCM, m is the mass of the NePCM under-relaxation was proposed by Voller and Swaminathan [83].
sample used in the DSC test. In fact, dH/dt is the heat flow rate through Many available phase change models assume that the melting and
the NePCM, and dT/dt is the scanning rate of the DSC test (i.e. tem- solidification processes follow an identical Cp (T) or H (T) relationship,
perature increase or decrease rate). which may lead to divergence between simulation and experimental
Since the obtained Ceff (T) curve by DSC is usually non-linear, it data. This is because the onset temperature of solidification is usually
needs to be processed using piecewise linear interpolation in numerical lower than the end temperature of melting, which is the so called phase
simulations. It is worth mentioning that the sample mass and DSC change hysteresis [84]. To consider this phenomenon in a consecutive
scanning rate can remarkably influence the obtained results in the dy- charging and discharging process (e.g. PCMs for passive cooling of
namic mode, in which the sample is heated or cooled at a constant rate. building envelopes), two Cp (T) or H (T) relationships should be used:
This is because the sample with a large size may not be in thermal one for charging and the other one for discharging [85].
equilibrium with the ambient environment under a fast heating/cooling Compared to the phase change hysteresis, the supercooling phe-
rate [71]. Iten et al. [72] suggested that the DSC scanning rate should nomenon is more difficult to be numerically simulated. Most numerical
be based on the practical cooling/heating rate of the selected applica- studies of supercooling were based on the enthalpy method. For ex-
tion. On the other hand, NePCM samples with a small mass (usually ample, Bony and Citherlet [84] proposed a supercooling model by using
5–50 mg) may not be entirely homogeneous, so the consistency be- two H (T) curves to separately model the discharging processes of a
tween the DSC results using different samples of the same material may PCM water tank with and without supercooling. In the H (T) curve for
not be guaranteed. To resolve this problem, the isothermal step mode describing supercooling, the authors assumed the crystallization pro-
can be used. cess of supercooled PCM is instantaneous and isenthalpic. The same
(2) Enthalpy method model was used by Huang et al. [74] to consider the supercooling of

6
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

hydrate salt based PCMs. Based on the EHCM, Jin et al. [86] developed that lies between the macroscale and the microscale. In the literature,
a supercooling model, which assumes there is an internal heat source in the most common mesoscale simulation technique is lattice-Boltzmann
the PCM to account for the latent heat released from the fast crystal- method (LBM). The LBM follows the discrete movement theory to
lization process. Similarly, Zhou and Han [87] developed a multiphase consider a set of fictive fluid particles confined to a regular lattice
model for supercooling by introducing a heat flux source term in the [102]. The particle streaming and collision are simulated by solving a
energy equation. It should be noted, however, that nanoparticles as a special discretized form of the continuous Boltzmann equation for the
nucleating agent can effectively mitigate the supercooling of PCM. Wu single-particle distribution function [103]. Different from the macro-
et al. [88], for instance, decreased the supercooling degree of water by scale techniques which directly discretize and solve the governing
70.9% by dispersing 0.2 wt% Al2O3 nanoparticles. In a study by Liu equations for continuum fluid mechanics (i.e. continuity, momentum,
et al. [89], 0.05 wt% graphene oxide nanosheets decreased the super- and energy), the LBM integrates the distribution functions of lattices to
cooling degree of water by 69.1% with advancing the onset time of recover the distribution of macroscopic properties of the fluid.
nucleation by 90.7%. Given these experimental results, most numerical There is a growing trend for using the LBM to solve solid–liquid
studies have neglected the supercooling of NePCMs in the solidification phase change problems in recent years, which can be classified into two
process [90]. main types: phase-field LBM [104] and enthalpy-based LBM [105]. The
In addition to the supercooling degree, many experiments [91,92] phase-field LBM requires extremely small grid spacing for tracking the
have proved that dispersing nanoparticles can increase the viscosity of solid–liquid interface, and thus it is computationally demanding. Con-
the base PCM, which may weaken the convection heat transfer in the sequently, the phase-field LBM is usually used to study the crystal
melting process. Hence, it is important to numerically investigate the growth rather than the flow and heat transfer characteristics of mate-
flow behavior of NePCMs. The key for a realistic flow behavior is to rials [106]. As for the enthalpy-based LBM, it is actually based on using
couple the physical states (i.e. solid, mushy, and liquid) with the ve- classical LB models (e.g. double distribution functions LB model [107]
locity values of computational cells. Particularly, the velocity should be and multi-relaxation-times LB model [108]) to solve the main part (i.e.
vanished in the solid cell. Three solid velocity correction methods are transient, convective and diffusion terms) of Eq. (6). Further, the latent
mainly used: heat source term in Eq. (6) is introduced as a source term in the tem-
(1) Switch-off method perature distribution function of LB model. The commonly used 2D and
It is a straightforward method that directly overwrites the velocity 3D discrete velocity models are D2Q9 and D3Q19, respectively. The
value to zero in the solid cells. Nevertheless, it may create convergence simulation of NePCMs using the LBM follows the same assumptions
issues due to the velocity discontinuity at the mushy zone [79]. To mentioned in Section 2.1.
avoid this, a ramped switch-off method can be used [93]. The LBM has the benefits of solving linear equations, simple pro-
(2) Variable viscosity method gramming, and handling complex boundary conditions. More im-
This method considers that the solid phase has a very large viscosity portantly, the LBM is very suitable and stable for parallel computing
value. For example, several numerical studies [94–97] defined the dy- owing to its explicit scheme, local interactions, and consequently a very
namic viscosity of the solid NePCM as 106 Pa·s. The ultra-high viscosity low communication-to-computation ratio [109]. However, there are
will increase the diffusive term in the momentum equation, thereby also some issues to be resolved. For example, a deviation term can be
inhibiting all motions in the solid cells. Different viscosity definitions produced in most LBM-based phase change models when recovering the
for the solid phase can be found in [98]. However, this method tends to corresponding macroscopic equations [110,111]. In addition, most of
produce divergency if the time step does not match with the grid size the NePCM models based on the LBM were not validated against a wide
[79]. range of experimental data, so there should be more investigations fo-
(3) Darcy source term method cused on this area.
This method considers the mushy zone as a porous medium in which
the convection flow is simulated following the Carman-Kozeny equa- 2.2.3. Molecular-scale simulation
tion. The combination of the enthalpy method and Darcy source term When the problem is investigated at the molecular-scale, the con-
method is the so called “enthalpy-porosity method [99]” (EPM), which tinuum assumption is no more valid. Molecular dynamic (MD) simu-
has been extensively used in the PCM/NePCM literature. In the EPM, a lation considers the interactions between a set of atoms and molecules
liquid fraction based Darcy source term S (β) is introduced in the mo- (system components) over a fixed period of time until the system
mentum equation to “switch on/off” the velocity in the case of melting/ reaches a dynamic evolution [102]. Since the time period is extremely
freezing, which can be given by: short, typical MD time steps are usually in the order of 1 fs (10−15 s).
The MD simulation solves Newton’s equations of motion to obtain the
(1 − β )2 trajectories of atoms and molecules, from which the physical quantity
S= Amush ui
(β 3 + ε ) (8) of interest can be derived. Further, molecular mechanics force fields or
interatomic potentials are used to calculate forces between the particles
where ε (usually is 0.001) is a small number to avoid dividing by
and their potential energies, which is the key for an accurate MD si-
zero, and Amush is the mushy zone constant.
mulation of the studied physical quantity [112].
The Amush represents the amplitude of the velocity damping; the
As mentioned in Section 2.2.1, the thermal characterizations of
larger the Amush, the faster the velocity is damped in the mushy zone.
NePCMs can be influenced by many factors, and the obtained results
Though Amush values ranging from 103 to 108 have been recommended
may have uncertainties. The MD simulation can be used to investigate
in most phase change studies, the use of different Amush values can exert
the thermophysical properties of NePCMs. Yu et al. [113] adopted an
significant influence on the simulated results. For example, a higher
optimized MD method (Fig. 4) to simulate the influence of SiO2 nano-
Amush vale (107) can lead to a delayed melting rate along with a more
particles on the thermophysical properties of NaCl molten salt PCM.
realistic evolution of melting front as compared to a lower one (105)
Their results showed that dispersing 2.4 vol% of SiO2 nanoparticles
[100]. Similar results were also obtained for the solidification processes
significantly improved the thermal conductivity of NaCl by up to
modeled with using different Amush vales [101]. Given the influence of
44.2%. The shear viscosity of molten NaCl was also decreased by up to
nanoparticles on the rheological behavior of the base PCM, the adopted
23.6%. Further, the melting point of NaCl increased with the volume
Amush in the numerical studies of NePCMs should be reported.
fraction of SiO2 nanoparticles. Zhang et al. [114] performed a non-
equilibrium MD to simulate ethylene–vinyl acetate paraffin PCMs dis-
2.2.2. Mesoscale simulation persed with different concentrations of graphene. The simulated results
Mesoscale simulations investigate the physical phenomena at a scale showed that the case with dispersing 0.7 wt% of graphene reached a

7
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

Fig. 4. Flow chart of the optimized MD simulation of NaCl-SiO2 molten salt NePCMs [113].

better thermal conductivity enhancement than those with dispersing By extending the Einstein’s equation [120], Brinkman [121] pro-
higher concentrations (1.5 wt%, 3.6 wt%, and 7.0 wt%). This trend may posed this model to predict the viscosity of particle suspensions with
be due to the complex interactions between ethylene–vinyl acetate and moderate loadings. The Brinkman model considers the dispersion of
graphene. Similar MD simulations of nanocarbons enhanced PCMs were one solute-molecule to a continuous solution. Several experiments
also conducted by Tafrishi et al. [115] and Babaei et al. [116]. Despite [122–124] have verified that the Brinkman model underestimated the
the thermal conductivity enhancement, the latent heat reduction viscosity of nanofluids, except in the case of low particle volume frac-
caused by nanoparticles should also be investigated. Zhao et al. [117], tions (ϕ < 3%).
for instance, indicated that dispersing 19.72 wt% CuO nanoparticles (2) Krieger-Dougherty (K-D) model
reduced the melting enthalpy of paraffin wax by as much as 51.5%. The −Aϕmax
significant latent heat reduction is due to the dense phase which results ϕ ⎞
μnp = μp ⎜⎛1 − ⎟
from the interactions between nanoparticles and PCM. ⎝ ϕmax ⎠ (10)
Table 2 summarizes the above mentioned numerical methods along
with their advantages and limitations. where ϕmax is the maximum packing factor, and A is the intrinsic
viscosity.
Krieger and Dougherty [125] derived this semi-empirical equation
2.3. Prediction models and factors based on the differential effective medium theory (DEMT). The values
of ϕmax and A for particles in the shapes of spheres, ellipsoids, cylinders,
As mentioned in Section 2.1, the effective thermal conductivity (knp) rods, and dumbbells can be found in [125,126]. However, comparisons
and viscosity (µnp) of NePCMs are very important for accurate simula- have shown that the K-D model can underestimate the viscosity of TiO2
tion results. Though the underlying mechanisms for the thermal and and Al2O3 nanofluids [127]. Chen et al. [128] pointed out that the best
rheological properties of nanofluids are still not well understood with prediction results can be achieved using the concentration of nano-
some remain controversial [119]. Several prediction models have been particle aggregates. For nanospheres, the aggregate size is about 3 times
extensively used in numerical studies of NePCMs to obtain the knp and of the primary nanoparticles size. Given these results, the K-D model
µnp. This section focusses on these models with their limitations and may not be very suitable for predicting the viscosity of nanoparticle
applicable ranges. suspensions.
Three empirical correlations Eqs. (11)–(13) were proposed based on
2.3.1. Effective dynamic viscosity experimental data, and they have been frequently adopted in the
(1) Brinkman model NePCM studies.
μp (3) Vajjha’s model
μnp =
(1 − ϕ)2.5 (9) μnp = 0.983e (12.959ϕ) μp , 0.01 ⩽ ϕ ⩽ 0.1 for Al2 O3 nanofluid (11)

8
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

is influenced by many factors, such as sample

A linear H (T) relationship may produce discrepancies between numerical


μnp = 0.919e (22.853ϕ) μp , 0.01 ⩽ ϕ ⩽ 0.06 for CuO nanofluid

demanding, which is related to the number of particles.


and isothermal phase change problem problems.
(12)

deviation term can be produced when recovering the corresponding


variable thermophysical properties is not straightforward
Vajjha [129] developed the equations based on the empirical cor-
relations by Sahoo et al. [130] and Namburu et al. [131]. Vajjha et al.
[132] suggested that Eqs. (11) and (12) are practical within 20–90 °C.
The authors used ethylene glycol/water (EG/W, 60:40) as the base

• Difficult to decide the most suitable force fields [113].


optimum under-relaxation factor is required [83].
fluid, but the average nanoparticle size was not reported.
(4) Corcione’s model
time step is needed for accuracy [66].

μp

viscosity is not a direct input [102].


μnp =

models for NePCMs are scare.


oscillation problems [79].

1 − 34.87(dn dp, e )−0.3ϕ1.03 (13)


efficiency [66].

computational efficiency [66].

equations [110,111].
• mass and DSC scanning rate [72].

where dn is the nanoparticle diameter, and dp,e is the equivalent


experimental results [80].

diameter of the base fluid molecule given by:


• A suitable C (T) relationship

1
• Difficult to solve sharp

3
⎛ 6M ⎞
dp, e = 0.1 ⎜
• ALowsmallcomputational

Nπρp,0 ⎟
(14)
• Computationally
⎝ ⎠
p

• Temperature

• Amacroscopic
where M is the molecular weight of the base fluid, N is the Avogadro
• Validated
• Modeling

number, and ρp,0 is the density of the base fluid at 20 °C.


Limitations

[102].

The above model was developed by fitting a large number of ex-


• Bulk
• Low
• and
• An

perimental data of nanofluids with a standard deviation of 1.84%, in-


cluding water, ethanol, EG, and propylene glycol (PG) dispersed with
TiO2, Al2O3, SiO2, and Cu nanoparticles [122]. The applicable ranges of
faster convergence rate under a Newton-Raphson procedure [82].
for both isothermal and smooth phase change problems.

for both isothermal and smooth phase change problems.

for non-equilibrium processes at the atomic level [118].

dn and ϕ were reported to be 25–200 nm and 0.0001–0.071, respec-


complex geometries and porous medium [103].

tively. Further, the correlation was independent of temperature within


• molecules [102]. for considering the interactions between
for smooth phase change problems (phase change

20–60 °C.
Table 3 summarizes the above prediction models of effective dy-
namic viscosity. Based on the reported models, there are several issues
simple for in-house programming.
equation is given in the general form.

to be resolved:
temperature range is greater than 2 °C [67]).

• Simulate dynamic properties of the system.


and simple programming

First, the base fluids studied in the above models are water, ethanol,
for simulating complex fluids.
parallel computing [109].

PG, and EG, whereas their rheological properties, molecular structures,


and intermolecular interactions are different from those of PCMs (e.g.
for complex geometries.
computational efficiency.

paraffin wax and eutectic salts). Using these nanofluid prediction


one unknown variable.

to reach convergence.

models to evaluate the viscosity of NePCMs may produce uncertainties.


Second, surfactants have been frequently used to improve the sus-
pension stability of nanoparticles, which was also found to increase the
• Mesoscopicfor nature
• Suitableequations

viscosity of nanofluids [133]. However, none of the above models


• Conceptually

considered this effect.


• Suitable for

Third, the morphology of nanoadditives has an important influence


• Suitable
• Suitable
• Suitable

• Suitable
• Suitable
• Suitable
Advantages

• Energy

• ALinear
• High
• Only

on the viscosity of nanofluids, but most of the above models are only
• Easy

applicable for microspheres or nanospheres. For example, a recent


study [134] have shown that nano-Al2O3 in the shapes of nanoparticles
and nanorods increased the viscosity of eutectic salt by 25% and 37%,
LAMMPS [115–117]; Materials Studio
COMSOL Multiphysics (FEM); ANSYS

COMSOL Multiphysics (FEM); ANSYS

COMSOL Multiphysics (FEM); ANSYS

respectively, However, dispersing multi-walled carbon nanotubes


(MWCNT) [135] and graphene [136] could reduce the viscosity of
nanofluids, which was attributed to the self-lubrication effect of
MWCNT and graphene. Hence, using the above models may produce
uncertainties for the NePCMs dispersed with non-spherical nano-
Common CFD codes
Summary of the main techniques for simulations of NePCMs.

OpenLB; Palabos

particles.
Fluent (FVM)

Fluent (FVM)

Fluent (FVM)

[113,114]

2.3.2. Effective thermal conductivity


The Maxwell model [137] based on the effective medium theory
(EMT) assumes there is no slip between the base fluid and particles, so it
is a “static” model. Hamilton and Crosser [138] further extended the
Maxwell model to non-spherical particles. However, the Hamilton-
Enthalpy-based
Phase change

Crosser (HC) model not only underestimated the thermal conductivity


enhancement by dispersing nanosized particles, but also failed to ad-
EHCM
model

HSTM

dress the influence of nanoparticle size and temperature dependency.


LBM

N.A.
EM

On the other hand, factors such as thermal dispersion, thermal interface


resistance, Brownian motion, and nanoparticle morphology also have
Simulation technique

significant impacts. Accordingly, several extensions were developed by


adding a “dynamic” part into the Maxwell-HC model [139].
(1) Thermal dispersion model
FVM/FEM
Table 2

kn + 2kp − 2ϕ (kp − kn )
kp + C′ (ρCp )np |u| ϕdn
LBM

knp =
MD

kn + 2kp + ϕ (kp − kn ) (15)

9
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

Table 3
Summary of prediction models of effective dynamic viscosity used in NePCM studies. Note: the values without a citation are summarized based on the reviewed
studies in Section 3 and 4.
Prediction model Factors Applicability

Morphology Particle size Temperature (°C) Concentration (vol.%)

Brinkman model ϕ, µp microsphere > 1000 nm [121] 0–170 °C 1–8%


K-D model ϕ, µp, ϕmax, A sphere, ellipsoid, cylinder, rod, dumbbell > 1000 nm [125] 20–40 °C 1–2%
Vajjha’s model ϕ, µp nanoparticle 10–100 nm 20–90 °C [129] 1–10% for Al2O3 [129]
1–6% for CuO [129]
Corcione’s model ϕ, µp, dn, dp,e nanoparticle 25–200 nm [122] 20–60 °C [122] 0.01–7.1% [122]

where u is the fluid velocity, and C’ is an empirical constant which is (4) Nan’s model
usually evaluated following the work by Wakao and Kaguei [140].
3 + ϕ [2β11 (1 − L11) + β33 (1 − L33 )]
It should be note, however, that the values of C’ were not explicitly knp = kp
3 − ϕ (2β11 L11 + β33 L33 ) (22)
stated in any NePCM studies except the studies [141,142]. The second
part of Eq. (15) represents the contribution of thermal dispersion effect, where
which is the so called dispersion thermal conductivity. The form of the
k11 − kp k33 − kp
dispersion thermal conductivity was proposed by Xuan and Roetzel β11 = , β33 =
[54], and it was adopted by Khanafer et al. [143] for modeling Cu kp + L11 (k11 − kp ) kp + L33 (k33 − kp ) (23)
nanofluids. However, the temperature dependency is not considered in where L11 and L33 are the geometrical factors of nanostructures, and
this model, which may produce deviations for simulation results. they can be expressed as:
(2) Vajjha’s model
a2 a
kn + 2kp − 2ϕ (kp − kn ) BT L11 = − cosh−1 a, for a > 1 and L33 = 1 − 2L11
knp = kp + 5 × 10 4γϕ (ρCp )p f (T , ϕ ) 2(a2 − 1) 2(a2 − 1)3 2
kn + 2kp + ϕ (kp − kn ) ρn dn
(24)
(16)
a2 a
where B is the Boltzmann constant (B = 1.3807 × 10−23 J/K), γ is L11 = + cos−1 a, for a < 1 and L33 = 1 − 2L11
2(a2 − 1) 2(1 − a2)3 2
the fraction of the liquid volume traveling with a nanoparticle and can
(25)
be defined as:
1
γ = 8.4407(100ϕ)−1.07304, 0.01 ⩽ ϕ ⩽ 0.1 for Al2 O3 nanofluid (17) L11 = L33 = , for a = 1
3 (26)
γ= 9.881(100ϕ)−0.9446, 0.01 ⩽ ϕ ⩽ 0.06 for CuO nanofluid (18) where a is the aspect ratio of nanostructures, k11 and k33 and are the
The empirical function f (T, ϕ) can be given by: equivalent thermal conductivities along different cell directions, and
they can be expressed as:
f (T , ϕ)
kn
T k ii = θLii kn
, i = 1, 3
= (2.8217 × 10−2ϕ + 3.917 × 10−3) + ( −3.0669 × 10−2ϕ 1+
Tref kp (27)
− 3.9112 × 10−3) (19) where
The second part of Eq. (16) captures the improved thermal con- Rkp

ductivity induced by Brownian motion of nanoparticles with con- θ=


⎧ 2+( 1
a ) ln
, for a ⩾ 1
⎨ (2 + 2a) Rkp
sidering the influence of temperature. The authors [144] suggested that , for a ⩽ 1
⎩ dn (28)
the applicable ranges of dn and T are 29–77 nm and 25–90 °C, respec-
tively. where R is the thermal boundary resistance between the nanos-
(3) Corcione’s model tructures and base fluid.
10 0.03 The main feature of this EMT based model is that the interfacial
⎡ T k ⎤
knp = kp ⎢1 + 4.4Re0.4Pr 0.66 ⎛⎜ ⎞⎟ ⎛⎜ n ⎞⎟ ϕ0.66⎥ thermal resistance and nanostructure morphology are considered
⎝ Tfr ⎠ ⎝ kp ⎠ (20) [146]. Hence, the Nan’s model is practical for nanostructures in the
⎣ ⎦
shapes of aligned continuous fibers, laminated flat plates, spheres, and
where Pr is the Prandtl number of the base fluid, and Tfr is the misoriented ellipsoidal particles. However, its accuracy is greatly af-
freezing point of the base fluid. Re is the nanoparticle Reynolds number fected by the assumption of R. For instance, the predicted thermal
which is defined as: conductivity of graphene nanoplatelets (GnP)/eicosane at 10 °C showed
2ρp BT the best agreement with experimental data for R of 9 × 10−9 m2·K/W
Re = [20]. At 20 °C, Harish et al. [147] validated the Nan’s model for GnP
πμp2 dn (21)
enhanced lauric acid with assuming the R to be 10−8 m2·K/W. On the
Corcione [122] developed this semi-empirical model by regression other hand, the Nan’s model has not been validated against the ex-
analysis. The model considers the nanoparticle size, fluid temperature, perimental data of liquid NePCMs, which requires further investiga-
and base fluid properties. It showed a standard deviation of 1.86% tions. Xu and Kleinstreuer [148] proposed a similar model which con-
against the experimental data of water, EG, and PG dispersed with Cu, siders the influence of nanoparticle aggregates.
Al2O3, CuO, and TiO2 nanoparticles. The applicable ranges of dn, ϕ, and Table 4 summarizes the above prediction models of effective
T were 10–150 nm, 0.002–0.09, and 21–51 °C, respectively. Chon et al. thermal conductivity. It is worth mentioning that in Eqs. (15) and (16),
[145] also developed a similar equation for Al2O3 nanofluids, which has the nanofluid velocity and temperature are used as correlation factors
a different definition of nanoparticle Reynolds number as compared to to control the dynamic terms. This is because the nanoparticle move-
Eq. (21). ment and its influence should be weakened in the solid. Some studies

10
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

Table 4
Summary of prediction models of effective thermal conductivity used in NePCM studies.
Prediction model Factors Applicability

Morphology Particle Size Temperature (°C) Concentration (vol.%)

Thermal dispersion model kn, kp, ϕ, C’, ρnp, Cp,np, u, dn nanoparticle 10–100 nm 0–80 °C 1–10%
Vajjha’s model kn, kp, ϕ, ρp, Cp, p, ρn, dn, T nanoparticle 29–77 nm [144] 25–90 °C [144] 1–10% [144]
Corcione’s model kn, kp, ϕ, ρp, Cp, p, µp, dn, T nanoparticle 10–150 nm [122] 21–51 °C [122] 0.2–9% [122]
Nan’s model kn, kp, ϕ, a, R, dn, ln fibers, plates, spheres, ellipsoidal particles Aspect ratio < 1000 20–70 °C 1–3%

Note: the values without a citation are summarized based on the reviewed studies in Section 3 and 4.

also introduced the liquid fraction as a correlation factor in the dynamic and-tube system based on the EPM. The K-D models showed that dis-
terms, which ensures these dynamic models can be smoothly reduced to persing 3 vol% of GnP increased the effective viscosity of NePCM by
the Maxwell model in the solid cells. On the other hand, using the above 36.1% and reduced the latent heat by 8.5%. The increased viscosity was
thermal conductivity models may produce the same issues facing the found to suppress the convection heat transfer, thereby greatly off-
viscosity models mentioned in Section 2.3.1. setting the benefit of the enhanced thermal conductivity of 205.8%.
Hence, the authors suggested to keep the GnP loading below 3 vol%. At
2.3.3. Other thermophysical properties a heat transfer fluid (HTF) temperature of 60 °C, dispersing 2 vol% of
Eq. (29)-(32) gives the prediction models of other effective ther- GnP significantly reduced the melting time by 41%. Following the same
mophysical properties used in the NePCM studies, including density (ρ), method, the authors [156,157] further compared the effects of different
specific heat (Cp), thermal expansion coefficient (β), and latent heat of nanocarbon morphologies and system orientations on the melting rate
fusion (L) [149]: of n-octadecane. 0D carbon nanodiamond (ND), 1D SWCNT, and 2D
GnP were used to enhance the PCM. A low particle loading of 1 vol%
ρnp = (1 − ϕ) ρp + ϕρn (29) was adopted in each case to maintain a moderate increase in the visc-
(ρCp )np = (1 − ϕ)(ρCp )p + ϕ (ρCp )n (30) osity of NePCM. As shown in Fig. 5, the melting time is shortened by
1%, 15%, and 25% for ND, SWCNT, and GnP in the vertical shell-and-
(ρβ )np = (1 − ϕ)(ρβ )p + ϕ (ρβ )n (31) tube system, respectively [156]. As for the horizontal system, the
melting time is reduced by 2%, 27%, and 40% for ND, SWCNT, and
(ρL)np = (1 − ϕ)(ρL)p (32) GnP, respectively [157]. Hence, the horizontal system is more efficient
Among the above properties, the most important one for the thermal in improving the effectiveness of carbon nanostructures.
performance of NePCMs is L. Since the nanoparticles do not change Kant et al. [158] investigated the melting of NePCMs in a square
phase, Eq. (32) considers the increase of nanoparticle volume fraction cavity. Their numerical method was based on the EHCM and Darcy
will reduce the latent heat of NePCMs. Nevertheless, this might not source term method. Three kinds of PCMs (capric acid, CaCl2·6H2O, and
always be true in experimental results. For example, Shaikh et al. [150] n-octadecane) were enhanced by 1 vol%, 3 vol%, and 5 vol% graphene
found that dispersing 1 vol% of carbon nanofibers (CNF), single-walled nanoparticles. The nanoparticles increased the heat conduction but
carbon nanotubes (SWCNT), and MWCNT increased the latent heat of considerably degraded the natural convection. The melting rate was
paraffin wax by 6.8%, 10.1%, and 13.0%, respectively. Zeng et al. improved in all cases, indicating the contribution of the enhanced
[151] and Wang et al. [152] also reported a minor increase in latent conduction heat transfer is large enough to make up the degraded
heat by adding MWCNT into pure PCMs. As for the nano metal oxides, convection heat transfer. It should be noted that the reduced latent heat
dispersing 0.3 wt% of TiO2 nanoparticles and CuO nanoparticles were is another important reason for the accelerated melting rate, particu-
found to enhance the latent heat of paraffin wax by as much as 15.7% larly in the case of a high particle loading. However, the authors did not
and 64.7%, reservedly [34]. Similar results were also obtained by dis- address this in both cases. A similar study was recently performed by
persing 2 wt% of α-Al2O3 nanoparticles into paraffin wax [153]. The Iachachene et al. [159], in which a trapezoidal cavity was used to store
reason for these results is related to the interactions between the mo- graphene nanoparticles enhanced paraffin wax. The thermal con-
lecules of base PCM and nanoparticles. If the potential of nanoparticle/ ductivity of paraffin wax was assumed to be enhanced by 20–100%. The
PCM molecular interaction is greater than that of PCM intermolecular simulation results showed that the increase of thermal conductivity
interaction, the latent heat may increase [150]. should be higher than 80% to achieve an improved heat transfer per-
Despite the latent heat of PCMs can be increased by nanoparticles, formance.
the vast majority of studies have observed degraded latent heat in both For GnP which is non-spherical, the prediction models discussed in
experiments [18,37,38] and MD simulations [117,154]. The incon- Section 2.3.2 may not be able to give results with acceptable accuracy.
sistent variations of latent heat may be one reason for the discrepancy Accordingly, Singh et al. [160] used Chu’s model [161] to evaluate the
between simulation and experimental results, since the numerical effective thermal conductivity of sugar alcohol/GnP with considering
models based on Eq. (32) always assume that the latent heat is reduced the GNP aspect ratio, interfacial thermal resistance, anisotropy, non-
by nanoparticles. Hence, more efforts should be devoted to an improved linearity effects, and GnP loading. Based on the EPM, a 2-D shell-and-
understanding of the underlying mechanisms for the variation of latent tube model was established to compare the effects of varying the fin
heat with nanoadditives. height and GnP loading on the melting rate. Fig. 6 shows that 5 vol% of
GnP can reduce the melting time by 54.8% which is higher than the
3. Melting studies of NePCMs optimized fin by 48.4%. Using both heat transfer enhancement tech-
niques, the melting time is notably shortened by 67.7%. As a follow-up
3.1. Nanocarbon enhanced PCMs to the study [160], Singh et al. [162] further demonstrated that the
degraded convection efficiency can be overcome using an optimized fin
Carbon nanomaterials, including graphite, carbon nanotubes (CNT), with dispersing 5 vol% GnP in a conical shell-and-tube container
and GnP, have been positively used to enhance pure PCMs due to their (Fig. 7). The charging time was considerably reduced by 57%. Further,
excellent thermal conductivity (2000–3500 W/m·K). Das et al. [155] a multi-attribute decision making (MADM) analysis was used to eval-
simulated the melting of n-eicosane/GnP in a vertically oriented shell- uate the material attributes of NePCMs. The results showed the thermal

11
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

Fig. 5. (a) 3D numerical model of the shell-and-tube system [156] and (b) total melting time for pure PCM and NePCMs with 1 vol% of nanocarbons in vertical and
horizontal systems [157].

conductivity has the highest normalized weight of 51%, followed by the 3.2. Nanometal enhanced PCMs
latent heat of 26%. The authors indicated that GnP is the best thermal
conductivity enhancer for PCMs, since it exhibits the highest thermal Nanometals are also well known for their remarkable thermal
conductivity enhancement along with the lowest latent heat reduction conductivity. Cu nanoparticles, with average thermal conductivity of
as compared to the nano metals and nano metal oxides. 400 W/m·K and diameter of 100 nm, has been widely used to enhance
Empirical viscosity models of eutectic PCMs (LiNO3-KCl) with dif- pure PCMs [164]. Darzi et al. [165] compared the benefits of dispersing
ferent GnP loadings were obtained by rotational rheometer [163]. Cu nanoparticles (diameter of 80 nm) and changing the shapes of tube
Fig. 8 shows that most prediction models, including the Brinkman heat exchanger (Fig. 9) for enhancing the melting rate of n-eicosane in a
model, have largely underestimated the real viscosity of GnP enhanced horizontally oriented cylindrical annulus. The inclusion of 2 vol% and
PCM. Using these prediction models may lead to a large under- 4 vol% of nanocopper shortened the charging time by 25% and 46%,
estimation of the degraded natural convection heat transfer in simu- respectively. Under the same operating condition, dispersing 4 vol% of
lated results. Their results showed that when using the measured nanocopper outweighed mounting 4 fins in terms of the charging rate.
viscosity, dispersing 1 vol% of GnP can even prolong the melting time However, dispersing nanoparticles did not eliminate the stable melting
by 6%, and dispersing 5 vol% of GnP only shortened the melting time pattern at the bottom region of the annulus, as shown in Fig. 10. This
by 17%. In the light of these results, it is necessary to use the measured can be addressed from two aspects. On one hand, the solid PCM next to
thermophysical properties to obtain credible numerical results. the heat exchanger will first melt to liquid phase at the beginning of
melting. The liquid PCM with a low thermal conductivity will wrap the

Fig. 6. Total melting time for different sets of heat transfer enhancement techniques [160].

12
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

empirical equation [168] considering the nanoparticle diameter, water


molecular size, and Brownian motion velocity. The enthalpy-based LBM
with a double distribution function (DDF) model was used to solve the
temperature and velocity fields. Mounting the heat exchanger at the
bottom section of the container with dispersing 4 vol% of nanocopper
was found to be the best heat transfer enhancement scheme. The en-
hancement is due to the shrinkage of the heat conduction dominated
region, so more regions benefit from the convection melting. However,
the increased viscosity of NePCM weakened the convection heat
transfer which is worthy of further attention. Using the same method,
similar conclusions were drawn in another two melting studies of
square cavity containers [169,170].
Feng et al. [107] developed a novel DDF-LBM involving density
evolution equation and temperature evolution equation. Their model
treated the latent heat source term with an implicit scheme, thereby
avoiding iteration steps in solving the temperature field. Results
showed that the Cu nanoparticle enhanced ice performed a higher heat
transfer rate in a square cavity heated from below. The time for
reaching a melting fraction of 0.7 was reduced by up to 17.2%. In-
creasing the volume fraction of Cu nanoparticles from 0 to 0.05 and
0.10, the energy storage rate was increased by 3.33% and 4.73% at a
Grashof number (Gr) of 5.0 × 104, and 1.5% and 5.33% at a Gr of
Fig. 7. Schematic of a conical shell-and-tube NePCM system with fins [160]. 2.5 × 105, respectively. Interestingly, in the case at high nanoparticle
loading and Gr, an asymmetric solid–liquid interface was formed
heat transfer tube, acting as a “thermal insulation” layer. The thickness (Fig. 11). This has been ascribed to the asymmetric distribution of the
of the insulation layer will grow with time, further inhibiting the heat convection cells.
from transferring to the outer region with solid PCM. This is the so Hosseinizadeh et al. [171] simulated the unconstrained melting of
called “self-insulating” effect of PCM [166]. On the other hand, the heat nanocopper/paraffin in a spherical container. The moving interface
transfer rate at the top region can be compensated by the convection between the air and molten NePCM was tracked by the volume of fluid
heat transfer, while the bottom region is usually dominated by the (VOF) model. The phase change rate under different Stefan numbers
conduction heat transfer. Dispersing nanoparticles only increased the (St, 0.05, 0.1 and 0.15) and nanoparticle loadings (0%, 2% and 4%)
PCM thermal conductivity, so it is unable to effectively mitigate the were examined. Since the mushy NePCM occupied most of the con-
imbalance of heat transfer mechanisms along the container height. To tainer, its thermal conductivity should be carefully considered by ex-
achieve a more uniform melting process, mounting fins at the bottom perimental measurements. Recently, Boukani et al. [172] further in-
region of the container with dispersing nanoparticles can be explored. vestigated the influence of adding nanocopper and changing the aspect
A similar study of annulus system was conducted by Jourabian et al. ratio of elliptical container on the unconstrained melting of n-octade-
[167] to discuss the combinations of changing the position of tube heat cane. The authors adopted the same numerical methods used in [171].
exchanger and Cu nanoparticle loading. The effective thermal con- For a given aspect ratio, adding nanocopper was reported to promote
ductivity of water/nanocopper PCM was determined by a semi- the melting rate and suppress the volume expansion of NePCM. The best
combination scheme was using an aspect ratio of 0.5 along with a

Fig. 8. Comparison between prediction models and experimental data [163].

13
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

Fig. 9. Cylindrical annulus with (a) circular tube, (b) vertical-oriented elliptical tube, (c) horizontal-oriented elliptical tube, and (d) finned circular tube [165].

charging time in all cases investigated. This is because the small


thermal conductivity enhancement by Al2O3 nanoparticles may not be
able to compensate the convection degradation. To resolve this down-
side, an idea is to prepare hybrid NePCMs by dispersing both metal and
metal oxide nanoparticles, which combines the thermal properties of
each nanoparticles.
Instead of using the prediction models which have great un-
certainties, Ghalambaz et al. [176] used experimental data to obtain the
evolution of NePCM thermophysical properties at different loadings of
Ag-MgO nanoparticles. The variation of solid–liquid interface in the
cases of four different dimensionless variables (k, µ) = (0,0), (5,18),
(18,18), (18,5) was particularly analyzed to find out the best compo-
sition scheme of the hybrid nanoparticles. The fastest melting rate was
achieved when the ratio of dimensionless k to µ was 18:5. The result
demonstrated that the melting rate can be improved only when the
Fig. 10. Stable and unstable melting patterns at the bottom and top region of
annulus [165].
thermal conductivity enhancement is much higher than the viscosity
augmentation. The same authors also simulated the melting of Ag-
MgO/water in a square cavity heated by an inner tube heat exchanger
nanoparticle loading of 3 vol%, reducing the melting time by 4.9%. [177]. The variation of solid–liquid interface was greatly affected by
A magnetic field can produce a volume force on the nanoparticles, the changed thermal conductivity rather than the changed viscosity.
thereby affecting the phase change process. Kohyani et al. [173] si- In recent years, high porosity metal and carbon foams with an ex-
mulated the melting process of cyclohexane/nanocopper in a square cellent thermal conductivity have also been widely used to enhance the
porous cavity with horizontal magnetic field. The effects of porosity, melting heat transfer of PCMs. However, using porous foams will in-
nanoparticle loading, Hartmann number (Ha, stands for the magnetic hibit the convection flow of liquid PCM more than mounting fins and
field intensity), and Rayleigh number (Ra) on the melting behavior dispersing nanoparticles [178]. In 2015, Hossain et al. [179] firstly
were considered individually. An implicit CFD code based on the EHCM reported on the melting of nano-CuO enhanced cyclohexane in an
was developed to solve the dimensionless governing equations. The aluminium foam. Their model (Fig. 12) considered the local thermal
strong magnetic field weakened the action of natural convection, equilibrium (LTE) assumption with ruling out the convection motion in
transforming the solid–liquid interface from a curvature to a straight the foam. Later, a hybrid nano-Al2O3/metal foam combination was
line. It was reported that changing the porosities of Al-porous medium examined by Mahdi and Nsofor [180], which considerably enhanced
was more effective in improving the melting rate than changing the the melting rate of paraffin RT82 in a triplex shell-and-tube system. The
nanoparticle loadings. However, the authors draw this conclusion authors modified the EPM formulations based on the local thermal non-
without considering the latent heat reduction. For example, changing equilibrium (LTNE) assumption to describe the heat and mass transfer
the porosity from 0.6 to 0.1 can at least reduce the total latent heat by of NePCM in a copper foam. The convection effect, non-Darcy effect,
50% since the porous foam does not change phase. This is an important Brownian motion, and temperature-dependent thermophysical proper-
reason for the greatly improved melting rate. This issue has been con- ties were considered. By dispersing nano-Al2O3 at loadings of 1 vol%,
sidered in a recent study by Li et al. [174], in which the porosity of 3 vol%, and 5 vol%, the melting time was shortened by 16.7%, 17.9%,
metal foam was 0.95 to keep a similar PCM mass as dispersing 5 vol% of and 19.7%, respectively. The melting in a foam with porosity of 0.95 is
Cu nanoparticles. much faster than that with porosity of 0.98 (18 min vs 120 min). The
authors explained this is because the total volume occupied by the
3.3. Nano metal oxide enhanced PCMs porous metal foam increased at a lower porosity, as well as the total
heat transfer area. Hence, the melting rate was notably improved with
Nano metal oxides such as Al2O3 and CuO nanoparticles has been only losing a small amount of latent heat.
another option to enhance pure PCMs. Most nano metal oxides have Optimizing the geometric structure to further improve the NePCM
relatively low thermal conductivity (about 40 W/m·K), so the loadings system was performed by Huo and Rao [108], in which a separate plate
of Al2O3 and CuO nanoparticles are usually higher than those of na- was mounted to mitigate the undesirable heat accumulation in the
nocarbons and nanometals. However, it would produce undesirable upper region of a square cavity. The nonuniform melting process results
viscosity augmentation and convection degradation. For example, from the imbalance of heat transfer mechanisms along the container
Parsazadeh and Duan [175] numerically investigated the melting rate height as analyzed before. For all Ra and nanoparticle loadings, a
of paraffin wax dispersed with Al2O3 nanoparticles in a finned shell- middle-located plate showed the fastest melting rate. Particularly, the
and-tube system. The results indicated that 4 vol% of Al2O3 nano- melting time was shortened by 10% for dimensionless plate height of
particles deteriorate the overall melting heat transfer, leading to longer

14
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

Fig. 11. Streamlines, temperature fields, and locations of solid–liquid interface at different dimensionless times (Gr = 2.5 × 105) [107].

Fig. 12. (a) Schematic of NePCM filled with a porous metal foam, (b) a magnified view of a representative control volume [179].

15
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

0.5 with dispersing 3 vol% of Al2O3 nanoparticles. The temperature


standard deviation was diminished by the separate plate, indicating a
mitigation of the heat accumulation in the upper region. However,
when the dimensionless plate height was less than 0.3, the heat accu-
mulation will slow down the overall melting rate. Similarly, Elbahjaoui
and El Qarnia [141] optimized the combinations of different para-
meters such as nano-Al2O3 loading, Re, Ra, and the aspect ratio of
NePCM slab. They reported that the melting rate was mainly affected by
Re and Ra rather than the nanoparticle loadings.
Among different PCM containers, shell-and-tube systems have been
extensively studied due to the ease of manufacture and modeling.
Varying The system inclination angle from 0 to 90°, a comparative
study by Pahamli et al. [181] showed that a vertically oriented shell-
and-tube system filled with paraffin RT50 and 4 vol% of CuO nano-
particles can achieve the fastest melting rate. Alomair et al. [94] ex-
perimentally and numerically investigated the influence of CuO nano-
particle loading and container height on the melting characteristics of
coconut oil. The NePCM dispersed with 0.25 vol% of CuO nanoparticles
was prepared by magnetic stirring and sonication, showing good sta-
bility without sedimentation. Both simulation and experimental results
showed an improved melting rate, but a small portion of gradually
diminished solid PCM/NePCM remained at the bottom region for a
prolonged period of time. This has been ascribed to the imbalance of
heat transfer mechanisms along the height of containers, particularly
the ones with a high aspect ratio. Similar results were also obtained in
the studies [182,183]. To achieve a uniform melting process, other heat
transfer enhancement techniques are needed to support the NePCM.
Using the same methodology, the melting behavior of nano-CuO/n-
octadecane in a horizontal cylinder and a square cavity were in-
vestigated by Dhaidan et al. [95,96]. Parsazadeh and Duan [184] per-
formed a statistical numerical study on both dispersing CuO nano-
particles in PCM and in HTF. The authors adopted the response surface
method (RSM) to analyze the effects of three parameters on the system
response, including the particle loading in PCM, particle loading in
HTF, and the HTF temperature. As shown in Fig. 13, two 3D plots were
used to elaborate the correlations between the above parameters. It is
clear that there is an optimum value for the nanoparticle loading in Fig. 13. 3D liquid fraction plots for (a) nanoparticle loading in PCM and HTF
PCM, which lies between 0.02 and 0.04. This is because dispersing temperature (b) nanoparticle loading in PCM and in HTF [184].
3.5 vol% of CuO nanoparticles has less negative impact on the natural
convection as compared to dispersing 7.0 vol% of CuO nanoparticles.
4. Solidification studies of NePCMs
The convection degradation is an important adverse effect induced
by nano metal oxides, particularly in the case of a high particle loading.
4.1. Nanocarbon enhanced PCMs
Arasu and Mujumdar [185] simulated the melting of nano-Al2O3 en-
hanced paraffin in a 25 mm × 25 mm cavity heated from different or-
Different from the melting process, the solidification of PCMs is
ientations. In all cases studied, dispersing 2 vol% nano-Al2O3 showed
predominated by conduction heat transfer throughout the entire pro-
better heat transfer performance than dispersing 5 vol% nano-Al2O3.
cess [187–189]. Hence, the negative effect viscosity augmentation and
Similar conclusions were also drawn by Arıcı et al. [186]. Given these
convection degradation may be eliminated in the solidification process
results, it is important to consider the coupling effects of the enhanced
of NePCMs. Another benefit is that nanoparticles can effectively miti-
conduction and the reduced convection. Dimensionless numbers such
gate the supercooling of NePCMs [190]. In 2015, Fan et al. [191] firstly
as the Nusselt number (Nu) and the Grashof number (Gr) can be used to
simulated the unidirectional solidification of GnP enhanced dodecanol.
evaluate the ratio of conduction to convection. As previously discussed,
Rather than using the prediction models, the authors experimentally
the viscosity augmentation is not likely to offset the melting enhance-
measured the effective thermophysical properties of NePCMs at dif-
ment in the case of a moderate nanoparticle loading. However, when
ferent GnP loadings. The HSTM was used to model the 1D phase change
the loading exceeds a certain value, the melting enhancement could be
(Stefan) problem. The relative solidification rates were quantitatively
outweighed. Numerical and experimental studies by Dhaidan et al. [97]
compared, showing an improvement of 34% by dispersing 1.0 wt% of
have verified that the melting rate of nano-CuO/n-octadecane was
GnP. The authors pointed out that the simulation is based on the EMT,
notably improved when increasing the nanoparticle loading from 0 to
so the migration of GnP nanoparticles and their interactions with the
1 wt%, whereas the enhancement ratio decreased when increasing the
moving solid–liquid interface, such as particle segregation and mor-
loading from 1 to 3 wt% and from 3 to 5 wt%. From a material point of
phological instability are not considered.
view, a high nanoparticle loading tends to cause particle agglomeration
Later, Singh et al. [192] numerically analyzed the solidification of
and segregation issues. Hence, further numerical and experimental
GnP/eutectic salt in a finned shell-and-tube system. The fin aspect ratio
studies are needed to investigate the optimum thermal properties of
has been optimized in the melting study [160]. The thermophysical
NePCMs.
properties of NePCM were treated as temperature-dependent poly-
Table 5 summarizes the above reviewed studies to which the readers
nomials in the EPM equations. The solidification time was reduced to
can refer.
7.6 h in the case of finned PCM, and it was further reduced to 5.3 h by

16
Table 5
Summary of numerical studies on the melting process of NePCMs.
Base PCM Nanostructure Concentration Melting Container Prediction model CFD code and method Amush Results and findings Ref.
T. Xiong, et al.

point

n-eicosane GnP with l of 1 μm and δ of 2 vol% 36 °C vertical shell-and- K-D model for µ Fluent based on EPM 105 41% melting time ↓ [155]
10 nm tube Nan’s model for k convection heat transfer was degraded due to the
increased µ
n-eicosane carbon ND with d of 10 nm 1 vol% 36 °C vertical shell-and- K-D model for µ Fluent based on EPM 105 1% melting time ↓ [156]
SWCNT with l of 1 μm and d tube Nan’s model for k 15% melting time ↓
of 1 nm 26% melting time ↓
GnP with l of 1 μm and δ of
10 nm
n-eicosane carbon ND with d of 10 nm 1 vol% 36 °C horizontal shell-and- K-D model for µ Fluent based on EPM 105 2% melting time ↓ [157]
SWCNT with l of μm and d of tube Nan’s model for k 27% melting time ↓
1 nm 40% melting time ↓
GnP with l of 1 μm and δ of
10 nm
capric acid graphene nanoparticles 1 vol% 32 °C square cavity Brinkman model for µ COMSOL based on EHCM 106 dispersing graphene increased the melting rate but [158]
CaCl2·6H2O 3 vol% 29 °C Vajjha’s model for k and Darth source term also degraded the convection heat transfer within
n-octadecane 5 vol% 28.2 °C method large cavities
paraffin wax graphene nanoparticles 0.5–2.0 vol% 46 °C trapezoidal cavity Brinkman model for µ Fluent based on EPM N.A. the enhancement of k should be higher than 80% to [159]
k was assumed to be achieve improved performance
enhanced by 20–100%
sugar alcohol GnP with l of 10 μm and δ of 1 vol% 166.3 °C vertical shell-and- Vajjha’s model for µ Fluent based on EPM 105 5% GnP outperformed optimized fin [160]
10 nm 3 vol% tube Chu’s model [161] for 67.7% melting time ↓ by 5% GnP + optimized fin
5 vol% k convection heat transfer was weakened due to the
increased µ and fin height
eutectic salt LiNO3- GnP with l of 10 μm and δ of 1 vol% 166 °C vertical conical Brinkman model for µ Fluent based on EPM 105 57% melting time ↓ by 5% GnP + optimized fin [162]

17
KCl 10 nm 3 vol% shell-and-tube Chu’s model [161] for
5 vol% k
eutectic salt LiNO3- GnP with l of 10 μm and δ of 1 vol% 166 °C vertical shell-and- Experimental data for µ Fluent based on EPM 105 the prediction models largely underestimated the µ [163]
KCl 10 nm 3 vol% tube Chu’s model [161] for of NePCM
5 vol% k 6% melting time ↑ by 1% GnP
43% melting time ↓ by 5% GnP + optimized fin
n-eicosane Cu nanoparticles with d of 2 vol% 36 °C horizontal shell-and- Brinkman model for µ Fluent based on EPM N.A. 25% melting time ↓ [165]
80 nm 4 vol% tube thermal dispersion 46% melting time ↓
model for k 4% GnP outperformed mounting 4 fins
convection heat transfer was weakened due to the
increased µ and fin height
water Cu nanoparticles with d of 2 vol% 0 °C horizontal shell-and- Brinkman model for µ enthalpy-based DDF-LBM N.A. 4% Cu nanoparticles + heat exchanger mounted at [167]
100 nm 4 vol% tube Patel’s model [168] for with D2Q9 lattice model the bottom achieved the best heat transfer
k enhancement
water Cu nanoparticles with d of 1 vol% 0 °C square cavity Brinkman model for µ enthalpy-based DDF-LBM N.A. 33% melting time ↓ by 2% Cu nanoparticles + heat [169]
100 nm 2 vol% Patel’s model [168] for with D2Q9 lattice model exchanger mounted at the bottom
k
water Cu nanoparticles with d of 1 vol% 0 °C square cavity Brinkman model for µ enthalpy-based DDF-LBM N.A. 50% melting time ↓ by 2% Cu [170]
100 nm 2 vol% Patel’s model [168] for with D2Q9 lattice model nanoparticles + optimized tube heat exchanger
k arrays
water Cu nanoparticles with d of 5 vol% 0 °C square cavity Brinkman model for µ enthalpy-based DDF LBM N.A. 17.9% melting time ↓ [107]
100 nm 10 vol% thermal dispersion with D2G9 and D2Q9 5.33% energy storage rate ↑
model for k lattice models
paraffin wax Cu nanoparticles 2 vol% 29 °C sphere Brinkman model for µ Fluent based on EPM with 105 12% melting time ↓ [171]
4 vol% thermal dispersion VOF model 25% melting time ↓
model for k
n-octadecane Cu nanoparticles with d of 1 vol% 28 °C elliptical capsule Brinkman model for µ Fluent based on EPM with 105 1.8% melting time ↓4.9% melting time ↓ [172]
80 nm 3 vol% Vajjha’s model for k VOF model
(continued on next page)
Applied Thermal Engineering 178 (2020) 115492
Table 5 (continued)

Base PCM Nanostructure Concentration Melting Container Prediction model CFD code and method Amush Results and findings Ref.
point
T. Xiong, et al.

cyclohexane Cu nanoparticles 1 vol% 7 °C square cavity Brinkman model for µ in-house CFD code based N.A. changing porosity outperformed changing [173]
5 vol% static Maxwell model on EHCM nanoparticle loading
for k a strong magnetic field lead to an increased melting
rate and a straight solid–liquid interface
paraffin wax Cu nanoparticles with d of 1 vol% 35 °C vertical shell-and- Vajjha’s models for µ Fluent based on EPM 105 25.9% melting time ↓ [174]
15 nm, 50 nm and 100 nm 5 vol% tube and k nanoparticles with small diameter showed a better
heat transfer enhancement than the ones with big
diameter
porous foam showed much better melting
enhancement than nanoparticles
paraffin wax Al2O3 nanoparticles with d of 4 vol% 47 °C Vertical shell-and- Vajjha’s models for µ Fluent based on EPM with 105 dispersing Al2O3 decreased the overall heat [175]
59 nm tube and k VOF model transfer rate, leading to longer charging time in all
cases
water Ag nanoparticles with d of 2 vol% 0 °C square cavity experimental data for µ MATLAB based on EPM 1.6 × 106 the best heat transfer enhancement was obtained [176]
25 nm and MgO nanoparticles and k when the ratio of dimensionless k to µ was 18:5
with d of 40 nm
water Ag nanoparticles with d of 5 vol% 0 °C square cavity experimental data for µ MATLAB based on EPM 1.6 × 106 the solid–liquid interface was greatly affected by [177]
25 nm and MgO nanoparticles and k the changed k rather than the changed µ.
with d of 40 nm
cyclohexane CuO nanoparticles 5–20 vol% 7 °C square cavity Brinkman model for µ NATURE based on EHCM N.A. porous foam showed much better melting [179]
static Maxwell model enhancement than nanoparticles
for k
paraffin wax Al2O3 nanoparticles 1 vol% 82 °C horizontal shell-and- Vajjha’s models for µ Fluent based on EPM with N.A. 19.7% melting time ↓ by 5% Al2O3 nanoparticles [180]
3 vol% tube and k considering local thermal porous foam showed much better melting
5 vol% equilibrium enhancement than nanoparticles

18
paraffin wax Al2O3 nanoparticles 3 vol% N.A. square cavity Vajjha’s models for µ enthalpy-based MRT-LBM N.A. 10% melting time ↓ [108]
5 vol% static Maxwell model with D2Q9 lattice model the separate plate should be mounted at a certain
for k height for a uniform melting process
paraffin wax Al2O3 nanoparticles 2–8 vol% 47 °C rectangle cavity Brinkman model for µ in-house CFD code based 1.6 × 106 by up to 14.25% melting time ↓ [141]
thermal dispersion on EHCM the Re and Ra showed higher influence than
model for k nanoparticle loadings
paraffin wax CuO nanoparticles 2 wt% 50 °C shell-and-tube with experimental data for µ in-house CFD code based 106 4.56% melting time ↓ [181]
4 wt% different and k on EPM 11.16% melting time ↓
orientations the vertical shell-and-tube system showed the
highest melting rate
coconut oil CuO nanoparticles 3 vol% 24 °C vertical shell-and- Brinkman model for µ COMSOL based on EHCM N.A. 11.1% melting time ↓ [94]
5 vol% tube static Maxwell model 13.8% melting time ↓
for k both simulation and experiment showed
accelerated melting rate
coconut oil CuO nanoparticles 3 vol% 24 °C vertical cylinder Brinkman model for µ COMSOL based on EHCM 105 the use of prediction models caused discrepancies [183]
5 vol% static Maxwell model and Darth source term between simulation and experiment
for k method
n-octadecane CuO nanoparticles 1 wt% 28 °C horizontal cylinder Brinkman model for µ COMSOL based on EHCM N.A. 4.5% melting time ↓ [95]
3 wt% thermal dispersion 10.1% melting time ↓
5 wt% model for k 12.8% melting time ↓
n-octadecane CuO nanoparticles 1 wt% 28 °C square cavity Brinkman model for µ COMSOL based on EHCM N.A. 5.5% melting time ↓ [96]
3 wt% thermal dispersion 8.9% melting time ↓
5 wt% model for k 11.7% melting time ↓
paraffin wax CuO nanoparticles with d of 2–4 vol% for HTF 55 °C vertical shell-and- Brinkman model for µ Fluent based on EPM 105 adding 3.5% CuO nanoparticles in PCM [184]
10 nm 3.5–7 vol% for tube Vajjha’s models for k outperformed adding 7% CuO in PCM
PCM
paraffin wax Al2O3 nanoparticles 2 vol% 47 °C square cavity Vajjha’s models for µ Fluent based on EPM N.A. adding 2% Al2O3 nanoparticles in PCM [185]
5 vol% and k outperformed adding 5% Al2O3 in PCM
(continued on next page)
Applied Thermal Engineering 178 (2020) 115492
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

[186]

[97]
Ref.

the melting enhancement rate was decreased with


outperformed adding 3% and 5% Al2O3 in PCM
adding 1% Al2O3 nanoparticles in PCM
Results and findings

the particle loading

Fig. 14. Schematic of the triplex shell-and-tube TES system with fins [160].

adding 5 vol% of GnP. Further, decreasing the St (HTF temperature)


and increasing the Re (HTF flow rate) also improved the solidification
Amush

N.A.
108

rate. Nevertheless, reducing the St is more efficient than increasing the


Re in enhancing the heat transfer of TES.
More recently, Alizadeh et al. [193] developed a novel triplex shell-
CFD code and method

Fluent based on EPM

Fluent based on EPM

and-tube system (Fig. 14) and expedited the freezing of water using V-
shaped fins and SWCNT. Their model was based on the HSTM and
solved by the standard Galerkin finite element method (SGFEM). The
fin design was optimized by the RSM with considering the angle be-
tween fin branches, the fin length, and the fin thickness. The fin length
and branch angle greatly affected the solidification rate, while the in-
fluence of the fin thickness was insignificant. The freezing time was
Brinkman model for µ
Vajjha’s models for µ

shortened by 33.1% by adding 5 vol% of SWCNT and 46.4% by


thermal dispersion
Prediction model

mounting fins, respectively. It should be noted that a part of the con-


tainer space is occupied by the fins, which will lead to a reduced latent
model for k

heat. This is also a reason for the faster freezing rate by mounting fins
and k

over adding SWCNT. However, this issue was not considered by the
authors.
Note: d, δ, l, ↓ and ↑ stand for diameter, thickness, length, reduction, and enhancement, respectively.
horizontal shell-and-
square cavity

4.2. Nanometal enhanced PCMs


Container

In 2007, Khodadadi and Hosseinizadeh [194] firstly reported on the


tube

accelerated freezing process of water enhanced by Cu nanoparticles


with an average diameter of 10 nm. Their method based on the EMT
and single-phase assumption was later used in a body of NePCM studies.
Melting
point

47 °C

28 °C

Results showed that the freezing time of water was shortened from
3000 s to 1400 s by dispersing 20 vol% of nanocopper. However, the
authors did not consider the particle sedimentation and dispersion
stability under such a high particle loading. Both experimental and
Concentration

numerical investigations are needed to ascertain the optimal particle


1–3 vol%

loading ranges, within which the NePCM can be considered as a single-


1 wt%
3 wt%
5 wt%

phase mixture. In addition, the largely reduced latent heat should also
be considered.
Al2O3 nanoparticles with d of

CuO nanoparticles with d of

Recently, more and more simulation-based studies have compared


different heat transfer enhancement techniques for promoting the so-
lidification rate of PCMs. For example, the freezing process of water in a
wavy cavity (Fig. 15) was studied by Abdollahzadeh and Esmaeilpour
Nanostructure

[195] to quantify the contributions of adjusting container surface wa-


viness and dispersing Cu nanoparticles. The authors maintained the
59 nm

same container volume for each case studied. Dispersing nanoparticles


9 nm

showed the best enhancement in Case II with a waviness of 0.2 m and a


Gr of 106, reducing the freezing time by up to 8.5%. Increasing the
Table 5 (continued)

surface waviness prolonged the freezing time whereas at a high Gr,


increasing the waviness could reduce the freezing time. The authors
n-octadecane
paraffin wax

stated that adjusting the surface waviness does not reduce the latent
Base PCM

heat, which is the advantage over dispersing nanoparticles. A similar


study can be found in [90].
Modularized PCM slab is another good option to increase the

19
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

Fig. 15. Schematic of the vertical wavy cavity filled with NePCM [195].

Fig. 16. (a) TES system consists of NePCM slabs and (b) computational domain [196].

contact surface area between PCM and HTF. Elbahjaoui and Qarnia dispersing nanoparticles is more efficient than mounting fins in the
[196] numerically studied the solidification of nano-copper/n-octade- melting process rather than in the solidification process. Similar con-
cane filled in a vertical slab with considering natural convection, as clusions were also emphasized in the comparison between using porous
shown in Fig. 16. The effects of varying the particle loading, slab aspect medium and dispersing nanoparticles [174].
ratio, and the dimensionless HTF temperature were examined. The It has been verified by numerical comparison that the influence of
solidification rate was improved with increasing the particle loading natural convection heat transfer can be ruled out in the solidification
and slab aspect ratio, and decreased with increasing the HTF tem- process [197]. This assumption makes the simulation process simpler,
perature. The best heat transfer enhancement scheme was by dispersing which has been widely adopted in recent studies. For example, a series
8 vol% of nanocopper under a slab aspect ratio of 6 and a dimensionless of numerical studies were performed by Lohrasbi et al. [198–201] to
HTF temperature of 1.1, reducing the solidification time by as much as model the freezing process of nanocopper/water in different finned
26.9%. shell-and-tube tanks, as shown in Fig. 17. In a typical analysis, the re-
The advantages of nanoparticles over other heat transfer enhance- search group firstly used the RSM to optimize the fin design, and then
ment techniques may be lost in the solidification process. For example, compared the effectiveness of mounting fins and dispersing nano-
the discharging time of a horizontal shell-and-tube system was reduced particles. In the study [44], dispersing 2.5 vol% and 5.0 vol% of Cu
by 9% and 16% by dispersing 2 vol% and 4 vol% of Cu nanoparticles nanoparticles improved the discharging rate by 1.099 and 1.208 time,
into n-eicosane [165]. Whereas mounting four fins shortened the dis- respectively. Whereas mounting fins improved the discharging rate by
charging time by 28%, which outperformed the enhancement by Cu as much as 5.749 time. Given the expedition of discharging rate and the
nanoparticles. Since dispersing 4 vol% of Cu nanoparticles out- reduction in latent heat, the authors concluded that mounting fins
performed mounting four fins in the melting process, it can be con- showed much better heat transfer enhancement than dispersing nano-
cluded that extending the heat transfer area is more efficient than particles.
dispersing nanoparticles in the solidification process. In the melting
process, the nanoparticles can better accommodate the convection flow,
while mounting fins can greatly suppress the convection flow. Hence,

20
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

Fig. 17. Finned PCM tanks studied in (a) [198], (b) [199,200], and (c) [201].

4.3. Nano metal oxide enhanced PCMs including adding nano-Al2O3 alone, using metal foam alone, and using
both metal foam and nano-Al2O3. The authors compared the LTE model
An interest study was recently carried out by Sheikholeslami and used in [179] with the LTNE model used in [180], and they verified that
Mahian [202], in which CuO nanoparticles and external magnetic field the simpler LTE model can obtain results with acceptable accuracy. This
were firstly used to improve the freezing process of water in a porous has been ascribed to the small temperature difference between the
annulus. The Koo-Kleinstreuer-Li (KKL) prediction model [203] was NePCM and metal foam during the heat conduction dominated solidi-
used for calculating the effective thermal conductivity and viscosity of fication process. Based on an LTE-modified EPM, the complete solidi-
the NePCM. The RSM was used to analyze the effects of three factors, fication time was shortened by 19.6% (8 vol% nanoparticle loading),
including the nanoparticle loading, Ha (magnetic force), and Ra. The 77.7% (0.98 porosity with 8 vol% nanoparticle loading), and 96.5%
complete freezing time of water was found to linearly decrease with the (0.95 porosity with 8 vol% nanoparticle loading), respectively.
nano-CuO loading and the Ha, and non-linearly increase with the Ra. More recently, Mahdi and Nsofor [206] improved the solidification
The freezing time was shortened by 14% for a nano-CuO loading of of paraffin RT82 by means of a hybrid strategy, including mounting fins
4 vol% and 23.5% for a Ha of 10, respectively. and adding nano-Al2O3. The combinations of using different fin arrays
Triplex shell-and-tube systems have been frequently used to extend mounted on the inner and outer tubes and dispersing nanoparticles with
the contact surface area between HTF and PCM, so they will show different loadings were analyzed, with considering the volume usage
better heat transfer rates than the normal shell-and-tube systems. (i.e. the ratio of occupied volume by nanostructures or fins to total PCM
Recently, a triplex shell-and-tube tank filled with nano Al2O3/paraffin volume) in each technique. Results indicated that the solidification
RT82 was studied by Mahdi and Nsofor [204]. User defined function process was significantly shortened using the hybrid strategy, by as
(UDF) codes were used to input the temperature-dependent thermo- much as 33.4%. However, for the same volume usage, a much better
physical properties of NePCM. The solidification process of NePCM at heat transfer efficiency can be achieved by mounting fins alone. The 2D
two HTF temperatures (65 °C and 70 °C) were investigated, respec- models established in [180,204–206] are shown in Fig. 18. These stu-
tively. Dispersing nano Al2O3 did not show much difference at the in- dies have given a benchmark of using CFD techniques to compare dif-
itial stage, but the solidification rate increased with the particle loading ferent heat transfer enhancement techniques and evaluate their in-
and elapse of time as heat conduction dominated the heat transfer re- dividual contributions to PCM systems.
gime. A total solidification time reduction of 20% was achieved by Table 6 summarizes the above reviewed studies to which the readers
adding 8 vol% of Al2O3 nanoparticles. can refer.
Later on, Mahdi and Nsofor [205] used open-celled porous copper
foam to further improve the discharging performance of the same
NePCM system studied in [180,204]. Three cases were simulated,

Fig. 18. 2D triplex shell-and-tube systems with: (a) NePCM [204], (b) fins with NePCM [206] and (c) porous foam with NePCM [180,205].

21
Table 6
Summary of numerical studies on the solidification process of NePCMs.
Base PCM Nanostructure Concentration Solidification point Container Prediction model CFD code and Amush Results and findings Ref.
T. Xiong, et al.

method

dodecanol GnP 0.5 wt% 24 °C heat transfer slab experimental data for in-house CFD N.A. 26% solidification time ↓ [191]
1.0 wt% k code based on 34% solidification time ↓
HSTM
eutectic salt LiNO3- GnP with l of 10 μm and δ of 1 vol% 166 °C vertical shell-and- Vajjha’s model for µ Fluent based on 105 29.1% solidification time ↓ [192]
KCl 10 nm 3 vol% tube Chu’s model [161] for EPM 39.8% solidification time ↓
5 vol% k 48.5% solidification time ↓
water SWCNT 2.5 vol% 0 °C vertical triple shell- static Maxwell model SGFEM based on N.A. 13.9% solidification time ↓ [193]
5.0 vol% and-tube for k HSTM 33.1% solidification time ↓
46.4% solidification time ↓ by mounting fins
water Cu nanoparticles with d of 10 vol% 0 °C square cavity Brinkman model for µ Fluent based on 105 33.3% solidification time ↓ [194]
10 nm 20 vol% thermal dispersion EPM 53.3% solidification time ↓
model for k
water Cu nanoparticles with d of 2.5 vol% 0 °C wavy cavity Brinkman model for µ in-house CFD N.A. by up to 8.5% solidification time ↓ [195]
1 nm 5.0 vol% thermal dispersion code based on
model for k EPM
water Cu nanoparticles with d of 5 vol% 0 °C wavy cavity Brinkman model for µ in-house CFD N.A. by up to 18.4% solidification time ↓ [90]
1 nm 10 vol% thermal dispersion code based on
model for k EPM
n-octadecane Cu nanoparticles 2 vol% 28 °C rectangle cavity Brinkman model for µ in-house CFD 1.6 × 106 by up to 26.9% solidification time ↓ [196]
8 vol% static Maxwell model code based on
for k EHCM
n-eicosane Cu nanoparticles with d of 2 vol% 36 °C horizontal shell- Brinkman model for µ Fluent based on N.A. 9% solidification time ↓ [165]
80 nm 4 vol% and-tube thermal dispersion EPM 16% solidification time ↓
model for k mounting fins outperformed dispersing nanoparticles in

22
all cases
water Cu nanoparticles 2.5 vol% 0 °C vertical shell-and- static Maxwell model SGFEM based on N.A. optimized fins showed much better solidification [201]
5.0 vol% tube for k HSTM enhancement than dispersing nanoparticles
paraffin wax Cu nanoparticles with d of 1 vol% 35 °C vertical shell-and- Vajjha’s models for µ Fluent based on 105 by up to 28.2% solidification time ↓ [174]
15 nm, 50 nm and 100 nm 5 vol% tube and k EPM nanoparticles with small diameter achieved a better
heat transfer enhancement than the ones with big
diameter
porous foam showed much better solidification
enhancement than dispersing nanoparticles
water Cu nanoparticles 2.5 vol% 0 °C vertical shell-and- static Maxwell model SGFEM based on N.A. 9.9% solidification time ↓ [200]
5.0 vol% tube for k HSTM 20.8% solidification time ↓
optimized fins showed much better solidification
enhancement than dispersing nanoparticles
water Cu nanoparticles 2.5 vol% 0 °C vertical shell-and- static Maxwell model SGFEM based on N.A. 8.8% solidification time ↓ [199]
5.0 vol% tube for k HSTM 16.9% solidification time ↓
optimized fins showed much better solidification
enhancement than dispersing nanoparticles
water Cu nanoparticles 2.5 vol% 0 °C vertical shell-and- static Maxwell model SGFEM based on N.A. optimized fins showed much better solidification [198]
5.0 vol% tube for k HSTM enhancement than dispersing nanoparticles
water CuO nanoparticles 4 vol% 0 °C porous annulus KKL model [203] for µ SGFEM based on N.A. 14% solidification time ↓ by 4% nano-CuO [202]
cavity and k HSTM 23.5% solidification time ↓ by a magnetic field with Ha
of 10
paraffin wax Al2O3 nanoparticles 3 vol% 82 °C horizontal triplex Vajjha’s models for µ Fluent based on 106 9.7% solidification time ↓ [204]
5 vol% shell-and-tube and k EPM 14% solidification time ↓
8 vol% 19.8% solidification time ↓
the solidification reduction percentage was almost
independent of the HTF temperature
(continued on next page)
Applied Thermal Engineering 178 (2020) 115492
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

5. Discussions

[205]

[206]
Ref.

5.1. Potential reasons for discrepancies

much better solidification enhancement than dispersing


77.7% solidification time ↓ by 8% nano-Al2O3 + 0.98

96.5% solidification time ↓ by 8% nano-Al2O3 + 0.95

under the same volume usage, mounting fins showed


Significant thermal enhancements by dispersing various nanos-

porous foam showed much better solidification


19.6% solidification time ↓ by 8% nano-Al2O3

tructures in PCMs have been simulated by most of the numerical studies

enhancement than dispersing nanoparticles


summarized in Table 5 and 6. Nevertheless, some experimental works
have reported that dispersing nanoparticles actually degraded the
overall heat transfer performance, particularly in the cases with strong
natural convection. For example, Ho and Gao [40] found the melting of
n-octadecane doped with 5 wt% and 10 wt% of nano Al2O3 was strongly
deteriorated in a square cavity. Zeng et al. [41] also demonstrated that
Results and findings

porous copper foam

porous copper foam

dispersing 1 wt% and 2 wt% of CNT greatly improved the melting rate
of dodeconal in a vertical cylinder at the initial stage. But as the melting
nanoparticles

proceeded the trend was reversed, and the dodeconal temperature


showed a much faster increase than the ones dispersed with CNT. As for
the solidification process, Fan and Khodadadi [42] reported that the
solidification time of nano-CuO/cyclohexane was linearly shortened
with the particle loading in the simulated results. However, the soli-
dification time was non-linearly shortened with the nanoparticle
Amush

106

106

loading in the experimental results. The potential reasons for these


discrepancies can be explained from three aspects:
(1) Unstable dispersion of the nanoparticles
Fluent based on

Fluent based on
CFD code and

All the experimental works [40–42] have observed particle sedi-


mentation issues at the end of the experiment, so it is difficult to prove
method

their repeatability. Since the flow velocity of the liquid NePCM is very
EPM

EPM

small in the natural convection, the velocity of nanoparticle sedi-


mentation should not be neglected in the cases without using stabilizers
Vajjha’s models for µ

Vajjha’s models for µ

or surfactants. This is an important reason for the discrepancy between


Prediction model

simulation and experimental results since any of the reported NePCM


models have not considered the particle sedimentation. On the other
hand, Alomair et al. [94] experimentally demonstrated that a NePCM
and k

and k

with very small particle sedimentation can achieve a faster melting rate
than the pure PCM. A better dispersion stability of nanoparticles also
results in stronger nanoparticle/PCM molecular interactions, thereby
horizontal triplex

horizontal triplex

improving the thermal properties of NePCM [150]. Hence, as long as


shell-and-tube

shell-and-tube

the dispersion stability of nanoparticles in the base PCM is satisfied, the


Container

NePCM can be considered as a homogeneous mixture and modeled


using single-phase approaches.
Several techniques can be used to improve the dispersion stability of
Note: d, δ, l, and ↓ stand for diameter, thickness, length, and reduction, respectively.
Solidification point

nanoparticles in the base PCM, including chemical treatment and


physical treatment. In the former technique, surfactants have been ex-
tensively used to stabilize the dispersed nanoparticles, either by elec-
trostatic approach [207] or surface modification approach [208]. In the
82 °C

82 °C

later technique, ultrasonication [209] and ball milling [210] have been
used to stabilize water-based and oil-based nanofluids. The colloidal
Concentration

stability of liquid NePCMs (nanofluids) can be mainly evaluated by


visual inspection [211], UV–vis spectrophotometer [212], and ζ-po-
3 vol%
5 vol%
8 vol%

1 vol%
2 vol%
3 vol%

tential method [213,214]. For example, Wensel et al. [211] improved


the suspension stability of SWCNT/Fe2O3/MgO in the water by adding
sodium dodecylbenzene sulfonate (SDBS) as the surfactant. The nano-
fluids were prepared using ultrasonication. As shown in Fig. 19, the
nanofluid with using surfactant (c) has a very uniform color distribution
Al2O3 nanoparticles

Al2O3 nanoparticles

than the one without using surfactant (b). Hwang et al. [212] used
UV–vis method to test the relationship between the absorbance of
Nanostructure

MWCNT, Fullerene, CuO/oil nanofluids and particle concentration. The


nanofluids were improved by adding sodium dodecyl sulfate (SDS) as
the surfactant. From the obtained relationship, the nano dispersion
stability can be estimated. As for the ζ potential method, it has been
used by Wang et al. [213] and Li et al. [214] to evaluate the dispersion
Table 6 (continued)

behavior of nanofluids improved by surfactants. The ζ-potential is


highly related to the absorbance of nanofluids, and a higher ζ-potential
paraffin wax

paraffin wax

represents a better dispersion stability of the system. On the other hand,


Base PCM

the dispersion stability of nanoparticles can influence the thermal sta-


bility of NePCMs. Hence, the NePCM thermal conductivity and latent
heat after many thermal cycles can also be used to evaluate the

23
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

conductivity. But when the particle loading exceeds a threshold, the


increased viscosity will strongly degrade the convection efficiency and
in turn slow down or even counteract the contribution of the improved
conduction. Hence, it is important to numerically analyze the optimum
ranges of particle loading for different PCM systems. In addition, the
nanostructures with high surface area to volume ratio and small dia-
meter show better heat transfer enhancements. Owning to the ad-
vantages of high thermal conductivity, high aspect ratio, large interface
contact area, and the minor side effects on latent heat reduction and
viscosity augmentation, GnP shows the best heat transfer enhancement
for PCM systems. It can be also found that most simulation-based stu-
dies have adopted the EPM due to its simplicity and efficiency in
modeling the convection melting of NePCMs. However, the mushy zone
constant should be carefully decided in the EPM, which requires further
attention.
The container geometries and orientations can also influence the
effectiveness of nanoparticles. For example, GnP enhanced n-eicosane
showed a faster melting rate in the horizontal shell-and-tube system
than the vertical one [156,157]. This is because the influence of natural
convection is not strong in the container with a low aspect ratio, so
Fig. 19. Color distributions of (a) water, (b) nanofluid without using surfactant, dispersing nanoparticles is more effective in this case. However, a re-
and (c) nanofluid with using surfactant [211]. verse finding was obtained in [181], showing that NePCM in the ver-
tical shell-and-tube system melted faster than that in the horizontal one.
By comparing the two studies, it was found that the HTF tube used in
dispersion stability of nanoparticles [215].
[181] is much smaller than the one used in [156,157], so the melting in
(2) Uncertainties of the prediction models
[181] is more affected by natural convection which is more efficient
Most of the currently used prediction models of effective viscosity
than adding nanoparticles in enhancing the melting process.
and thermal conductivity have uncertainties for NePCMs. This can also
In comparison with mounting fins and using porous metal foams,
produce uncertainties to the simulations involving conduction and
dispersing nanostructures can better accommodate the convection flow
convection heat transfer. Particularly, the Brinkman model has been
of the liquid NePCM. But the overall enhancement by nanoparticles is
reported to largely underestimate the real increase in the effective
still moderate as compared to the fins and metal foams. Another pro-
viscosity of PCM dispersed with non-spherical GnP [163] and CNT [41].
blem is that nanoparticles can only enhance the overall conduction heat
It has also been pointed out that simulated results using different pre-
transfer, so it is less efficient in mitigating the heat accumulation pro-
diction models can show a deviation, which is because some models did
blem induced by the strong natural convection in the upper region of
not consider the influence of nanoparticle diameter [216]. Based on
containers with a high aspect ratio [94,108,165]. To further improve
these results, it is suggested to use the measured viscosity and thermal
the system performance, hybrid strategies by means of dispersing na-
conductivity in the numerical simulation.
noparticles and extending heat transfer area are suggested.
(3) Limitations of the numerical assumptions
The vast majority of NePCM studies are based on single-phase ap-
proaches assuming the NePCM is a stable and homogeneous mixture,
5.2.2. Solidification process of NePCM
and thus some physical phenomenon were not considered in any of the
Based on the reported simulation results [155–157,174,191,192],
reported studies. For example, the prediction models mentioned in
nanoparticles was found to be more efficient in the solidification pro-
Section 2.3.2 did not consider the influence of nanoparticle migration
cess than the melting process. This is because nanoparticles will cause
induced by Brownian motion and thermophoresis. Nanoparticles in the
viscosity augmentation and consequently offset or even degrade the
liquid PCM may aggregate at a high particle loading, which is due to the
improved conduction heat transfer in the melting process. However, the
large surface area to volume ratio and high surface energy [217]. This
solidification process is predominated by conduction heat transfer, so
may lead to heterogeneous distribution of thermophysical properties
the side effect of viscosity augmentation can be eliminated. A clear
throughout the control volume. Further, the dynamic diffusion of na-
trend can be also identified that the solidification rate always increases
noparticles and their interactions with the moving solid–liquid interface
with the nanoparticle loading. But at high particle loadings, the single-
can cause engulfment and/or rejection of nanoparticles [142]. Given
phase assumption may break down due to the nanoparticle sedi-
the above issues, the single-phase assumption may break down. Ac-
mentation and migration issues. Accordingly, experimental works are
cordingly, additional considerations are necessary in the classical phase
needed to ascertain the nanoparticle loading ranges within which the
change formulations. For instance, a recent numerical study by Hasadi
single-phase assumption is reasonable to be used. Likewise, dispersing
[218] considered the mass transport of Cu nanoparticles in the melting
GnP with high aspect ratio and excellent thermal conductivity shows
process of water, showing a concentration field induced by the nano-
the best solidification enhancement as compared to the nanometals and
particle rejection at the solid–liquid interface.
nano metal oxides.
As for the solidification process, the effectiveness of dispersing na-
5.2. Basic phase change behavior of NePCM noparticles is clearly inferior to that of using fins and metal foams. This
is decided by the material properties and limitations of nanostructures,
5.2.1. Melting process of NePCM as well as the main heat transfer mechanism in the solidification pro-
Based on the reported simulation results [158–160], conduction and cess. Hence, dispersing nanoparticles is actually more suitable for the
convection are two main heat transfer mechanisms, and the convection scenarios which only need a complete melting/freezing cycle rather
is more efficient in improving the melting process. Hence, a majority of than a very fast melting/freezing rate. For example, for PCM integrated
numerical studies have investigated the impact of nanoparticles on the walls and roofs which have a strict U-value requirement, nanoparticles
melting rate. There is a clear trend that the melting rate accelerates can improve the effective utilization of latent heat without increasing
with the particle loading, which is due to the enhanced thermal the system weight and cost.

24
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

5.3. Research gaps and future outlook (1) The melting rate of NePCMs can be improved by nanoparticles
within a certain range of loading. The nanoparticles will also cause
Several research gaps and outlook for further work have been viscosity augmentation and consequently offset or even degrade the
identified from the reported studies: improved conduction heat transfer. Numerically investigate the
optimal nanoparticle properties and loadings for given systems is
(1) Since the current prediction models have large uncertainties which significant to improve the phase change process and heat storage
can lead to discrepancy between simulation and experiment rate.
[163,183], there is an urgent need to develop new empirical models (2) The solidification rate of NePCMs always increases with the nano-
for NePCMs. For this purpose, experimental studies using different particle loading. Under high particle loadings, the particle ag-
characterization methods are strongly encouraged. It is also im- gregation and sedimentation issues should be considered in the si-
portant to understand the diffusion of nanostructures and its in- mulation since the single-phase assumption may not be valid. The
fluence on the thermophysical properties of solid, mushy, and li- reduced latent heat by nanoparticles also improves the heat release
quid NePCMs. Potential approaches include MD simulation rate.
[113–117] and particle image velocimetry (PIV) [219]. (3) The effectiveness of nanoparticles is more prominent in the solidi-
(2) Most of the current prediction models have not considered the fication process than in the melting process. The effectiveness can
temperature dependency of effective viscosity, which also has not also be influenced by the container structures, geometries and or-
been considered in most of the reported simulations. In fact, the ientations, which is related to the material properties and main heat
viscosity of NePCM was found to decrease with temperature in transfer mechanisms.
some experiments [92,220]. Hence, the degraded convection heat (4) The selection of suitable nanostructures should be based on the
transfer may be recovered with the increase of temperature. It is thermal conductivity, aspect ratio, interface contact area, and the
suggested to consider the viscosity-temperature dependency in the side effects on latent heat reduction and viscosity augmentation.
numerical model. Nanocarbons, particularly GnP and CNT, showed the best heat
(3) Most of the reported studies have used the melting/solidification transfer enhancement for PCM systems.
rates to evaluate the TES performance of NePCM systems. However, (5) There is an urgent need to develop new empirical models for the
nanoparticles will usually reduce the latent heat as discussed in thermophysical properties of NePCMs, particularly thermal con-
Section 2.3.3. This is also a reason for the accelerated phase change ductivity and viscosity models. The uncertainties of the current
process. From another point of view, the increased melting/solidi- prediction models is the main reason for the discrepancy between
fication rate may not represent an improved heat storage/release simulation and experiment. Hence, it is suggested to use the mea-
rate [45]. Most studies did not consider this issue except [174] and sured thermal properties with considering the temperature de-
[206]. Hence, trade-off studies should be made to consider the in- pendency in simulations.
fluence of conduction enhancement, convection degradation, and (6) For NePCMs, trade-off studies must be conducted to consider the
latent heat reduction. The impact of dispersing nanoparticles on the influence of conduction enhancement, convection degradation, and
energy storage/release rate and exergy efficiency of system is worth latent heat reduction.
further study.
(4) Though the simulations based on single-phase approaches remain Declaration of Competing Interest
dominant in NePCM studies, they are unable to consider the in-
fluence of microscopic transport of nanoparticles such as the par- None.
ticle agglomeration and sedimentation issues. In this regard, new
two-phase and multi-phase models should be developed for Acknowledgements
NePCMs with considering the influence of nanoparticle diffusion.
For example, the LBM models with mesoscopic nature can be used This work has been supported by City Developments Limited (CDL)
in this area. But as discussed in Section 2.1, it is always difficult to project R-296-000-174-720 and NUS-AGC Inc. project R-296-000-183-
couple the physical states of the base PCM with the dynamics of 597. Acknowledgement is also given to Department of Building,
nanoparticles. National University of Singapore.
(5) To our knowledge, there are only a few studies [95,96,185,186]
available in which the numerical models have been validated Appendix A. Supplementary material
against the experimental data of phase change behavior of NePCMs.
The majority of the reported models have only been validated Supplementary data to this article can be found online at https://
against the previous experimental data of pure PCMs. This may not doi.org/10.1016/j.applthermaleng.2020.115492.
be a proper validation method. In the future, more experimental
studies of NePCMs are needed to validate the newly proposed nu- References
merical methods and models.
(6) Most of the reported studies (Table 5 and 6) have used 105 and 106 [1] I.E. Agency, World Energy Outlook (2019) 2019.
as the mushy zone constant in their NePCM models. However, these [2] L.H.J. Raijmakers, D.L. Danilov, R.A. Eichel, P.H.L. Notten, A review on various
Amush values also have not been validated by experiments. Since the temperature-indication methods for Li-ion batteries, Appl. Energy 240 (2019)
918–945.
liquid NePCM can become mudlike at a high nanoparticle loading [3] I.E. Agency, Global CO2 emissions in 2019, in, 2020.
[211], the Amush of NePCM and its influence should be reconsidered. [4] Y. Wang, X. Yang, T. Xiong, W. Li, K.W. Shah, Performance evaluation approach
Apparently, more investigations should be focused on this topic. for solar heat storage systems using phase change material, Energy Build. 155
(2017) 115–127.
[5] X. Yang, T. Xiong, J.L. Dong, W.X. Li, Y. Wang, Investigation of the Dynamic
6. Conclusions Melting Process in a Thermal Energy Storage Unit Using a Helical Coil Heat
Exchanger, Energies 10 (2017) 1129.
[6] H. Akeiber, P. Nejat, M.Z.A. Majid, M.A. Wahid, F. Jomehzadeh, I. Zeynali
Dispersing thermally conductive nanostructures offer an approach Famileh, J.K. Calautit, B.R. Hughes, S.A. Zaki, A review on phase change material
to remarkably enhance the thermal performance of pure PCMs, thereby (PCM) for sustainable passive cooling in building envelopes, Renew. Sustain.
improving their effectiveness in TES applications. Here, we provide a Energy Rev. 60 (2016) 1470–1497.
[7] P. Singh, S. Khanna, V. Becerra, S. Newar, V. Sharma, T.K. Mallick, D. Hutchinson,
comprehensive review on the simulation-based studies of NePCMs.
J. Radulovic, R. Khusainov, Power improvement of finned solar photovoltaic
Important conclusions are as follows:

25
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

phase change material system, Energy 193 (2020) 116735. nanoparticle-enhanced phase change material (NPCM) on solar still productivity,
[8] S. Hekmat, G.R. Molaeimanesh, Hybrid thermal management of a Li-ion battery J. Clean. Prod., 192 (2018) 9-29.
module with phase change material and cooling water pipes: An experimental [35] N. Şahan, M. Fois, H. Paksoy, Improving thermal conductivity phase change
investigation, Appl. Therm. Eng. 166 (2020) 114759. materials—A study of paraffin nanomagnetite composites, Sol. Energy Mater. Sol.
[9] Z. Huang, N. Xie, X. Zheng, X. Gao, X. Fang, Y. Fang, Z. Zhang, Experimental and Cells 137 (2015) 61–67.
numerical study on thermal performance of Wood's alloy/expanded graphite [36] X. Fang, L.-W. Fan, Q. Ding, X.-L. Yao, Y.-Y. Wu, J.-F. Hou, X. Wang, Z.-T. Yu, G.-
composite phase change material for temperature control of electronic devices, H. Cheng, Y.-C. Hu, Thermal energy storage performance of paraffin-based com-
Int. J. Therm. Sci. 135 (2019) 375–385. posite phase change materials filled with hexagonal boron nitride nanosheets,
[10] N.R. Jankowski, F.P. McCluskey, A review of phase change materials for vehicle Energy Convers. Manage. 80 (2014) 103–109.
component thermal buffering, Appl. Energy 113 (2014) 1525–1561. [37] J.M. Munyalo, X. Zhang, Particle size effect on thermophysical properties of na-
[11] C. Prieto, L.F. Cabeza, Thermal energy storage (TES) with phase change materials nofluid and nanofluid based phase change materials: A review, J. Mol. Liq. 265
(PCM) in solar power plants (CSP). Concept and plant performance, Appl. Energy (2018) 77–87.
254 (2019) 113646. [38] K.Y. Leong, M.R. Abdul Rahman, B.A. Gurunathan, Nano-enhanced phase change
[12] W. Kostowski, K. Pajączek, A. Pociecha, J. Kalina, P. Niedzielski, A. Przybył, materials: A review of thermo-physical properties, applications and challenges, J.
Methods of waste heat recovery – A compressor station case study, Energy Storage Mater. 21 (2019) 18–31.
Convers. Manage. 197 (2019) 111837. [39] R. Du, W. Li, T. Xiong, X. Yang, Y. Wang, K.W. Shah, Numerical investigation on
[13] S.-F. Li, Z.-H. Liu, X.-J. Wang, A comprehensive review on positive cold energy the melting of nanoparticle-enhanced PCM in latent heat energy storage unit with
storage technologies and applications in air conditioning with phase change ma- spiral coil heat exchanger, Build. Simul. 12 (2019) 869–879.
terials, Appl. Energy 255 (2019) 113667. [40] C.J. Ho, J.Y. Gao, An experimental study on melting heat transfer of paraffin
[14] J. Jaguemont, N. Omar, P. Van den Bossche, J. Mierlo, Phase-change materials dispersed with Al2O3 nanoparticles in a vertical enclosure, Int. J. Heat Mass
(PCM) for automotive applications: A review, Appl. Therm. Eng. 132 (2018) Transf. 62 (2013) 2–8.
308–320. [41] Y. Zeng, L.-W. Fan, Y.-Q. Xiao, Z.-T. Yu, K.-F. Cen, An experimental investigation
[15] J. Pereira da Cunha, P. Eames, Thermal energy storage for low and medium of melting of nanoparticle-enhanced phase change materials (NePCMs) in a
temperature applications using phase change materials – A review, Appl. Energy bottom-heated vertical cylindrical cavity, Int. J. Heat Mass Transf. 66 (2013)
177 (2016) 227–238. 111–117.
[16] J. Chen, S. Kang, J.E, Z. Huang, K. Wei, B. Zhang, H. Zhu, Y. Deng, F. Zhang, G. [42] L. Fan, J.M. Khodadadi, A theoretical and experimental investigation of uni-
Liao, Effects of different phase change material thermal management strategies on directional freezing of nanoparticle-enhanced phase change materials, J. Heat
the cooling performance of the power lithium ion batteries: A review, Journal of Transf. 134 (2012).
Power Sources, 442 (2019) 227228. [43] L.-W. Fan, Z.-Q. Zhu, Y. Zeng, Q. Lu, Z.-T. Yu, Heat transfer during melting of
[17] K.W. Shah, A review on enhancement of phase change materials - A nanomaterials graphene-based composite phase change materials heated from below, Int. J. Heat
perspective, Energy Build. 175 (2018) 57–68. Mass Transf. 79 (2014) 94–104.
[18] M.A. Kibria, M.R. Anisur, M.H. Mahfuz, R. Saidur, I.H.S.C. Metselaar, A review on [44] L.-W. Fan, Z.-Q. Zhu, Y. Zeng, Q. Ding, M.-J. Liu, Unconstrained melting heat
thermophysical properties of nanoparticle dispersed phase change materials, transfer in a spherical container revisited in the presence of nano-enhanced phase
Energy Convers. Manage. 95 (2015) 69–89. change materials (NePCM), Int. J. Heat Mass Transf. 95 (2016) 1057–1069.
[19] L.-W. Fan, X. Fang, X. Wang, Y. Zeng, Y.-Q. Xiao, Z.-T. Yu, X. Xu, Y.-C. Hu, K.- [45] N. Hu, Z.-Q. Zhu, Z.-R. Li, J. Tu, L.-W. Fan, Close-contact melting heat transfer on a
F. Cen, Effects of various carbon nanofillers on the thermal conductivity and en- heated horizontal plate: Revisited in the presence of nano-enhanced phase change
ergy storage properties of paraffin-based nanocomposite phase change materials, materials (NePCM), Int. J. Heat Mass Transf. 124 (2018) 794–799.
Appl. Energy 110 (2013) 163–172. [46] N.S. Dhaidan, Nanostructures assisted melting of phase change materials in var-
[20] X. Fang, L.-W. Fan, Q. Ding, X. Wang, X.-L. Yao, J.-F. Hou, Z.-T. Yu, G.-H. Cheng, ious cavities, Appl. Therm. Eng. 111 (2017) 193–212.
Y.-C. Hu, K.-F. Cen, Increased thermal conductivity of eicosane-based composite [47] Z. Ma, W. Lin, M.I. Sohel, Nano-enhanced phase change materials for improved
phase change materials in the presence of graphene nanoplatelets, Energy Fuels 27 building performance, Renew. Sustain. Energy Rev. 58 (2016) 1256–1268.
(2013) 4041–4047. [48] A. Kasaeian, L. bahrami, F. Pourfayaz, E. Khodabandeh, W.-M. Yan, Experimental
[21] S. Kim, L.T. Drzal, High latent heat storage and high thermal conductive phase studies on the applications of PCMs and nano-PCMs in buildings: a critical review,
change materials using exfoliated graphite nanoplatelets, Sol. Energy Mater. Sol. Energy Build., 154 (2017) 96-112.
Cells 93 (2009) 136–142. [49] H.W. Chiam, W.H. Azmi, N.M. Adam, M.K.A.M. Ariffin, Numerical study of na-
[22] M. Li, A nano-graphite/paraffin phase change material with high thermal con- nofluid heat transfer for different tube geometries – A comprehensive review on
ductivity, Appl. Energy 106 (2013) 25–30. performance, Int. Commun. Heat Mass Transfer 86 (2017) 60–70.
[23] A. Karaipekli, A. Biçer, A. Sarı, V.V. Tyagi, Thermal characteristics of expanded [50] A. Albojamal, K. Vafai, Analysis of single phase, discrete and mixture models, in
perlite/paraffin composite phase change material with enhanced thermal con- predicting nanofluid transport, Int. J. Heat Mass Transf. 114 (2017) 225–237.
ductivity using carbon nanotubes, Energy Convers. Manage. 134 (2017) 373–381. [51] L. Godson, B. Raja, D. Mohan Lal, S. Wongwises, Enhancement of heat transfer
[24] W. Cui, Y. Yuan, L. Sun, X. Cao, X. Yang, Experimental studies on the supercooling using nanofluids—an overview, Renew. Sustain. Energy Rev. 14 (2010) 629–641.
and melting/freezing characteristics of nano-copper/sodium acetate trihydrate [52] M.S. Mojarrad, A. Keshavarz, A. Shokouhi, Nanofluids thermal behavior analysis
composite phase change materials, Renew. Energy 99 (2016) 1029–1037. using a new dispersion model along with single-phase, Heat Mass Transf. 49
[25] J.L. Zeng, Z. Cao, D.W. Yang, L.X. Sun, L. Zhang, Thermal conductivity en- (2013) 1333–1343.
hancement of Ag nanowires on an organic phase change material, J. Therm. Anal. [53] S. Özerinç, A.G. Yazıcıoğlu, S. Kakaç, Numerical analysis of laminar forced con-
Calorim. 101 (2009) 385–389. vection with temperature-dependent thermal conductivity of nanofluids and
[26] J.-L. Zeng, F.-R. Zhu, S.-B. Yu, L. Zhu, Z. Cao, L.-X. Sun, G.-R. Deng, W.-P. Yan, thermal dispersion, Int. J. Therm. Sci. 62 (2012) 138–148.
L. Zhang, Effects of copper nanowires on the properties of an organic phase change [54] Y. Xuan, W. Roetzel, Conceptions for heat transfer correlation of nanofluids, Int. J.
material, Sol. Energy Mater. Sol. Cells 105 (2012) 174–178. Heat Mass Transf. 43 (2000) 3701–3707.
[27] T. Sreethawong, K.W. Shah, S.-Y. Zhang, E. Ye, S.H. Lim, U. Maheswaran, [55] S. Kumar, S.K. Prasad, J. Banerjee, Analysis of flow and thermal field in nanofluid
W.Y. Mao, M.-Y. Han, Optimized production of copper nanostructures with high using a single phase thermal dispersion model, Appl. Math. Model. 34 (2010)
yields for efficient use as thermal conductivity-enhancing PCM dopant, J. Mater. 573–592.
Chem. A 2 (2014). [56] Y. Ding, D. Wen, Particle migration in a flow of nanoparticle suspensions, Powder
[28] R. Parameshwaran, R. Jayavel, S. Kalaiselvam, Study on thermal properties of Technol. 149 (2005) 84–92.
organic ester phase-change material embedded with silver nanoparticles, J. [57] A. Behzadmehr, M. Saffar-Avval, N. Galanis, Prediction of turbulent forced con-
Therm. Anal. Calorim. 114 (2013) 845–858. vection of a nanofluid in a tube with uniform heat flux using a two phase ap-
[29] Y. Deng, J. Li, T. Qian, W. Guan, Y. Li, X. Yin, Thermal conductivity enhancement proach, Int. J. Heat Fluid Flow 28 (2007) 211–219.
of polyethylene glycol/expanded vermiculite shape-stabilized composite phase [58] A. Malvandi, M.R. Safaei, M.H. Kaffash, D.D. Ganji, MHD mixed convection in a
change materials with silver nanowire for thermal energy storage, Chem. Eng. J. vertical annulus filled with Al2O3–water nanofluid considering nanoparticle mi-
295 (2016) 427–435. gration, J. Magn. Magn. Mater. 382 (2015) 296–306.
[30] P. Chandrasekaran, M. Cheralathan, V. Kumaresan, R. Velraj, Enhanced heat [59] S.M. Vanaki, P. Ganesan, H.A. Mohammed, Numerical study of convective heat
transfer characteristics of water based copper oxide nanofluid PCM (phase change transfer of nanofluids: a review, Renew. Sustain. Energy Rev. 54 (2016)
material) in a spherical capsule during solidification for energy efficient cool 1212–1239.
thermal storage system, Energy 72 (2014) 636–642. [60] G. Liang, I. Mudawar, Review of single-phase and two-phase nanofluid heat
[31] M. Nourani, N. Hamdami, J. Keramat, A. Moheb, M. Shahedi, Thermal behavior of transfer in macro-channels and micro-channels, Int. J. Heat Mass Transf. 136
paraffin-nano-Al2O3 stabilized by sodium stearoyl lactylate as a stable phase (2019) 324–354.
change material with high thermal conductivity, Renew. Energy 88 (2016) [61] O. Mahian, L. Kolsi, M. Amani, P. Estellé, G. Ahmadi, C. Kleinstreuer,
474–482. J.S. Marshall, M. Siavashi, R.A. Taylor, H. Niazmand, S. Wongwises, T. Hayat,
[32] R.K. Sharma, P. Ganesan, V.V. Tyagi, H.S.C. Metselaar, S.C. Sandaran, Thermal A. Kolanjiyil, A. Kasaeian, I. Pop, Recent advances in modeling and simulation of
properties and heat storage analysis of palmitic acid-TiO2 composite as nano-en- nanofluid flows-Part I: Fundamentals and theory, Phys. Rep. 790 (2019) 1–48.
hanced organic phase change material (NEOPCM), Appl. Therm. Eng. 99 (2016) [62] Fluent 19.0 User's Guide, in, 2019.
1254–1262. [63] N.S. Bondareva, B. Buonomo, O. Manca, M.A. Sheremet, Heat transfer perfor-
[33] Q. He, S. Wang, M. Tong, Y. Liu, Experimental study on thermophysical properties mance of the finned nano-enhanced phase change material system under the in-
of nanofluids as phase-change material (PCM) in low temperature cool storage, clination influence, Int. J. Heat Mass Transf. 135 (2019) 1063–1072.
Energy Convers. Manage. 64 (2012) 199–205. [64] M. Arıcı, E. Tutuncu, A. Campo, Numerical investigation of melting of paraffin
[34] D. Dsilva Winfred Rufuss, L. Suganthi, S. Iniyan, P.A. Davies, Effects of wax dispersed with CuO nanoparticles inside a square enclosure, Heat Transf. Res.

26
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

49 (2018). numerical study of constrained melting of n-octadecane with CuO nanoparticle


[65] Y. Khattari, T. El Rhafiki, N. Choab, T. Kousksou, M. Alaphilippe, Y. Zeraouli, dispersions in a horizontal cylindrical capsule subjected to a constant heat flux,
Apparent heat capacity method to investigate heat transfer in a composite phase Int. J. Heat Mass Transf. 67 (2013) 523–534.
change material, J. Storage Mater. 28 (2020) 101239. [96] N.S. Dhaidan, J.M. Khodadadi, T.A. Al-Hattab, S.M. Al-Mashat, Experimental and
[66] S.N. Al-Saadi, Z. Zhai, Modeling phase change materials embedded in building numerical investigation of melting of phase change material/nanoparticle sus-
enclosure: a review, Renew. Sustain. Energy Rev. 21 (2013) 659–673. pensions in a square container subjected to a constant heat flux, Int. J. Heat Mass
[67] X. Jin, H. Hu, X. Shi, X. Zhou, X. Zhang, Comparison of two numerical heat Transf. 66 (2013) 672–683.
transfer models for phase change material board, Appl. Therm. Eng. 128 (2018) [97] N.S. Dhaidan, J.M. Khodadadi, T.A. Al-Hattab, S.M. Al-Mashat, Experimental and
1331–1339. numerical investigation of melting of NePCM inside an annular container under a
[68] Z. Andrássy, Z. Szánthó, Thermal behaviour of materials in interrupted phase constant heat flux including the effect of eccentricity, Int. J. Heat Mass Transf. 67
change, J. Therm. Anal. Calorim. 138 (2019) 3915–3924. (2013) 455–468.
[69] Y. Zhang, K. Du, J.P. He, L. Yang, Y.J. Li, Impact Factors Analysis of the Enthalpy [98] J.H. Nazzi Ehms, R. De Césaro Oliveski, L.A. Oliveira Rocha, C. Biserni, M. Garai,
Method and the Effective Heat Capacity Method on the Transient Nonlinear Heat Fixed Grid Numerical Models for Solidification and Melting of Phase Change
Transfer in Phase Change Materials (PCMs), Numer. Heat Transf., Part A: Appl. 65 Materials (PCMs), Appl. Sci. 9 (2019) 4334.
(2014) 66–83. [99] A.D. Brent, V.R. Voller, K.J. Reid, Enthalpy-porosity technique for modeling
[70] Y. Zhang, Modified computational methods using effective heat capacity model for convection-diffusion phase change: application to the melting of a pure metal,
the thermal evaluation of PCM outfitted walls, Int. Commun. Heat Mass Transf. Numer. Heat Transf. 13 (1988) 297–318.
108 (2019) 104278. [100] M. Fadl, P.C. Eames, Numerical investigation of the influence of mushy zone
[71] E. Günther, S. Hiebler, H. Mehling, R. Redlich, Enthalpy of Phase Change Materials parameter Amush on heat transfer characteristics in vertically and horizontally
as a Function of Temperature: Required Accuracy and Suitable Measurement oriented thermal energy storage systems, Appl. Therm. Eng. 151 (2019) 90–99.
Methods, Int. J. Thermophys. 30 (2009) 1257–1269. [101] M. Hameter, H. Walter, Influence of the Mushy Zone Constant on the Numerical
[72] M. Iten, S. Liu, A. Shukla, P.D. Silva, Investigating the impact of C p -T values Simulation of the Melting and Solidification Process of Phase Change Materials, in:
determined by DSC on the PCM-CFD model, Appl. Therm. Eng. 117 (2017) 65–75. Z. Kravanja, M. Bogataj (Eds.), Computer Aided Chemical Engineering, Vol. 38
[73] M. Iten, S. Liu, A. Shukla, Experimental validation of an air-PCM storage unit Elsevier, 2016, pp. 439–444.
comparing the effective heat capacity and enthalpy methods through CFD simu- [102] O. Mahian, L. Kolsi, M. Amani, P. Estellé, G. Ahmadi, C. Kleinstreuer,
lations, Energy 155 (2018) 495–503. J.S. Marshall, R.A. Taylor, E. Abu-Nada, S. Rashidi, H. Niazmand, S. Wongwises,
[74] H. Huang, Y. Xiao, J. Lin, T. Zhou, Y. Liu, Q. Zhao, Improvement of the efficiency T. Hayat, A. Kasaeian, I. Pop, Recent advances in modeling and simulation of
of solar thermal energy storage systems by cascading a PCM unit with a water nanofluid flows—Part II: Applications, Phys. Rep. 791 (2019) 1–59.
tank, J. Clean. Prod. 245 (2020) 118864. [103] Y.-L. He, Q. Liu, Q. Li, W.-Q. Tao, Lattice Boltzmann methods for single-phase and
[75] Z. Yinping, J. Yi, A simple method, the -history method, of determining the heat of solid-liquid phase-change heat transfer in porous media: A review, Int. J. Heat
fusion, specific heat and thermal conductivity of phase-change materials, Meas. Mass Transf. 129 (2019) 160–197.
Sci. Technol. 10 (1999) 201–205. [104] I. Rasin, W. Miller, S. Succi, Phase-field lattice kinetic scheme for the numerical
[76] Y. Qu, S. Wang, D. Zhou, Y. Tian, Experimental study on thermal conductivity of simulation of dendritic growth, Phys. Rev. E, Stat., Nonlinear, Soft Matter Phys. 72
paraffin-based shape-stabilized phase change material with hybrid carbon nano- (2005) 066705.
additives, Renew. Energy 146 (2020) 2637–2645. [105] W.-S. Jiaung, J.-R. Ho, C.-P. Kuo, Lattice Boltzmann method for the heat con-
[77] C. Yadav, R. Rekha Sahoo, Experimental analysis for optimum thermal perfor- duction problem with phase change, Numer. Heat Transf., Part B: Fundam. 39
mance and thermophysical parameters of MWCNT based capric acid PCM by using (2001) 167–187.
T-history method, Powder Technol. 364 (2020) 392–403. [106] R. Rojas, T. Takaki, M. Ohno, A phase-field-lattice Boltzmann method for mod-
[78] K.P. Venkitaraj, S. Suresh, Effects of Al2O3, CuO and TiO2 nanoparticles son eling motion and growth of a dendrite for binary alloy solidification in the pre-
thermal, phase transition and crystallization properties of solid-solid phase change sence of melt convection, J. Comput. Phys. 298 (2015) 29–40.
material, Mech. Mater. 128 (2019) 64–88. [107] Y. Feng, H. Li, L. Li, L. Bu, T. Wang, Numerical investigation on the melting of
[79] Z. Ma, Y. Zhang, Solid velocity correction schemes for a temperature transforming nanoparticle-enhanced phase change materials (NEPCM) in a bottom-heated rec-
model for convection phase change, Int. J. Numer. Meth. Heat Fluid Flow 16 tangular cavity using lattice Boltzmann method, Int. J. Heat Mass Transf. 81
(2006) 204–225. (2015) 415–425.
[80] B.L. Gowreesunker, S.A. Tassou, M. Kolokotroni, Improved simulation of phase [108] Y. Huo, Z. Rao, Lattice Boltzmann investigation on phase change of nanoparticle-
change processes in applications where conduction is the dominant heat transfer enhanced phase change material in a cavity with separate plate, Energy Convers.
mode, Energy Build. 47 (2012) 353–359. Manage. 154 (2017) 420–429.
[81] V.R. Voller, Implicit Finite—difference Solutions of the Enthalpy Formulation of [109] Q. Li, K.H. Luo, Q.J. Kang, Y.L. He, Q. Chen, Q. Liu, Lattice Boltzmann methods for
Stefan Problems, IMA J. Numer. Anal. 5 (1985) 201–214. multiphase flow and phase-change heat transfer, Prog. Energy Combust. Sci. 52
[82] J.C. Álvarez-Hostos, E.A. Gutierrez-Zambrano, J.C. Salazar-Bove, E.S. Puchi- (2016) 62–105.
Cabrera, A.D. Bencomo, Solving heat conduction problems with phase-change [110] R. Huang, H. Wu, Phase interface effects in the total enthalpy-based lattice
under the heat source term approach and the element-free Galerkin formulation, Boltzmann model for solid–liquid phase change, J. Comput. Phys. 294 (2015)
Int. Commun. Heat Mass Transf. 108 (2019) 104321. 346–362.
[83] V.R. Voller, C.R. Swaminathan, ERAL source-based method for solidification phase [111] R. Huang, H. Wu, A modified multiple-relaxation-time lattice Boltzmann model for
change, Numer. Heat Transf., Part B: Fundam. 19 (1991) 175–189. convection–diffusion equation, J. Comput. Phys. 274 (2014) 50–63.
[84] J. Bony, S. Citherlet, Numerical model and experimental validation of heat storage [112] S.M. Rassoulinejad-Mousavi, Y. Zhang, Interatomic potentials transferability for
with phase change materials, Energy Build. 39 (2007) 1065–1072. molecular simulations: a comparative study for platinum, gold and silver, Scient.
[85] (!!! INVALID CITATION !!! [72, 85]). Rep. 8 (2018) 2424–12410.
[86] X. Jin, H. Hu, X. Shi, X. Zhou, L. Yang, Y. Yin, X. Zhang, A new heat transfer model [113] Y. Yu, Y. Tao, Y.-L. He, Molecular dynamics simulation of thermophysical prop-
of phase change material based on energy asymmetry, Appl. Energy 212 (2018) erties of NaCl-SiO2 based molten salt composite phase change materials, Appl.
1409–1416. Therm. Eng. 166 (2020) 114628.
[87] G. Zhou, Y. Han, Numerical simulation on thermal characteristics of supercooled [114] M. Zhang, C. Wang, A. Luo, Z. Liu, X. Zhang, Molecular dynamics simulation on
salt hydrate PCM for energy storage: Multiphase model, Appl. Therm. Eng. 125 thermophysics of paraffin/EVA/graphene nanocomposites as phase change ma-
(2017) 145–152. terials, Appl. Therm. Eng. 166 (2020) 114639.
[88] S. Wu, D. Zhu, X. Li, H. Li, J. Lei, Thermal energy storage behavior of Al2O3–H2O [115] H. Tafrishi, S. Sadeghzadeh, R. Ahmadi, F. Molaei, F. Yousefi, H. Hassanloo,
nanofluids, Thermochim. Acta 483 (2009) 73–77. Investigation of tetracosane thermal transport in presence of graphene and carbon
[89] Y. Liu, X. Li, P. Hu, G. Hu, Study on the supercooling degree and nucleation be- nanotube fillers––A molecular dynamics study, J. Storage Mater. 29 (2020)
havior of water-based graphene oxide nanofluids PCM, Int. J. Refrig. 50 (2015) 101321.
80–86. [116] H. Babaei, P. Keblinski, J.M. Khodadadi, Thermal conductivity enhancement of
[90] S. Kashani, A.A. Ranjbar, M. Abdollahzadeh, S. Sebti, Solidification of nano-en- paraffins by increasing the alignment of molecules through adding CNT/graphene,
hanced phase change material (NEPCM) in a wavy cavity, Heat Mass Transf. 48 Int. J. Heat Mass Transf. 58 (2013) 209–216.
(2012) 1155–1166. [117] C.Y. Zhao, Y.B. Tao, Y.S. Yu, Molecular dynamics simulation of nanoparticle effect
[91] S. Motahar, N. Nikkam, A.A. Alemrajabi, R. Khodabandeh, M.S. Toprak, on melting enthalpy of paraffin phase change material, Int. J. Heat Mass Transf.
M. Muhammed, A novel phase change material containing mesoporous silica na- 150 (2020) 119382.
noparticles for thermal storage: A study on thermal conductivity and viscosity, Int. [118] F. Jabbari, A. Rajabpour, S. Saedodin, Thermal conductivity and viscosity of na-
Commun. Heat Mass Transf. 56 (2014) 114–120. nofluids: A review of recent molecular dynamics studies, Chem. Eng. Sci. 174
[92] B. Águila V, D.A. Vasco, P. Galvez P, P.A. Zapata, Effect of temperature and CuO- (2017) 67–81.
nanoparticle concentration on the thermal conductivity and viscosity of an organic [119] J.M. Khodadadi, L. Fan, H. Babaei, Thermal conductivity enhancement of nanos-
phase-change material, Int. J. Heat Mass Transf., 120 (2018) 1009-1019. tructure-based colloidal suspensions utilized as phase change materials for thermal
[93] Y. Dutil, D.R. Rousse, N.B. Salah, S. Lassue, L. Zalewski, A review on phase-change energy storage: A review, Renew. Sustain. Energy Rev. 24 (2013) 418–444.
materials: Mathematical modeling and simulations, Renew. Sustain. Energy Rev. [120] A. Einstein, Eine neue Bestimmung der Moleku¨ldimensionen, Ann. Phys. 19
15 (2011) 112–130. (1906) 289–306.
[94] M. Alomair, Y. Alomair, S. Tasnim, S. Mahmud, H. Abdullah, Analyses of Bio- [121] H.C. Brinkman, The Viscosity of Concentrated Suspensions and Solutions, J. Chem.
Based Nano-PCM filled concentric cylindrical energy storage system in vertical Phys., 20 (1952) 571-.
orientation, J. Storage Mater. 20 (2018) 380–394. [122] M. Corcione, Empirical correlating equations for predicting the effective thermal
[95] N.S. Dhaidan, J.M. Khodadadi, T.A. Al-Hattab, S.M. Al-Mashat, Experimental and conductivity and dynamic viscosity of nanofluids, Energy Convers. Manage. 52

27
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

(2011) 789–793. [154] Z. Rao, S. Wang, F. Peng, Molecular dynamics simulations of nano-encapsulated
[123] S.E.B. Maïga, S.J. Palm, C.T. Nguyen, G. Roy, N. Galanis, Heat transfer enhance- and nanoparticle-enhanced thermal energy storage phase change materials, Int. J.
ment by using nanofluids in forced convection flows, Int. J. Heat Fluid Flow 26 Heat Mass Transf. 66 (2013) 575–584.
(2005) 530–546. [155] N. Das, Y. Takata, M. Kohno, S. Harish, Melting of graphene based phase change
[124] C.T. Nguyen, F. Desgranges, G. Roy, N. Galanis, T. Maré, S. Boucher, H. Angue nanocomposites in vertical latent heat thermal energy storage unit, Appl. Therm.
Mintsa, Temperature and particle-size dependent viscosity data for water-based Eng. 107 (2016) 101–113.
nanofluids – Hysteresis phenomenon, Int. J. Heat Fluid Flow 28 (2007) [156] N. Das, Y. Takata, M. Kohno, S. Harish, Effect of carbon nano inclusion di-
1492–1506. mensionality on the melting of phase change nanocomposites in vertical shell-tube
[125] T.J.D.I.M. Krieger, A Mechanism for Non-Newtonian Flow in Suspensions of Rigid thermal energy storage unit, Int. J. Heat Mass Transf. 113 (2017) 423–431.
Spheres, Trans. Soc. Rheol. 3 (1959) 137–152. [157] N. Das, M. Kohno, Y. Takata, D.V. Patil, S. Harish, Enhanced melting behavior of
[126] I. Santamaria-Holek, C.I. Mendoza, The rheology of concentrated suspensions of carbon based phase change nanocomposites in horizontally oriented latent heat
arbitrarily-shaped particles, J. Colloid Interface Sci. 346 (2010) 118–126. thermal energy storage system, Appl. Therm. Eng. 125 (2017) 880–890.
[127] S.M.S. Murshed, K.C. Leong, C. Yang, Thermophysical and electrokinetic proper- [158] K. Kant, A. Shukla, A. Sharma, P. Henry Biwole, Heat transfer study of phase
ties of nanofluids – A critical review, Appl. Therm. Eng. 28 (2008) 2109–2125. change materials with graphene nano particle for thermal energy storage, Sol.
[128] H. Chen, Y. Ding, C. Tan, Rheological behaviour of nanofluids, New J. Phys. 9 Energy 146 (2017) 453–463.
(2007) 367. [159] F. Iachachene, Z. Haddad, H.F. Oztop, E. Abu-Nada, Melting of phase change
[129] R.S. Vajjha, Measurements of Thermophysical Properties of Nanofluids and materials in a trapezoidal cavity: Orientation and nanoparticles effects, J. Mol. Liq.
Computation of Heat Transfer Characteristics, in: Mechanical Engineering 292 (2019) 110592.
Department, Vol. M.S., University of Alaska Fairbanks, Fairbanks, Alaska, 2009. [160] R.P. Singh, S.C. Kaushik, D. Rakshit, Melting phenomenon in a finned thermal
[130] B.C. Sahoo, R.S. Vajjha, R. Ganguli, G.A. Chukwu, D.K. Das, Det ermination of storage system with graphene nano-plates for medium temperature applications,
Rheological Behavior of Aluminum Oxide Nanofluid and Development of New Energy Convers. Manage. 163 (2018) 86–99.
Viscosity Correlations, Pet. Sci. Technol. 27 (2009) 1757–1770. [161] K. Chu, C.-C. Jia, W.-S. Li, Effective thermal conductivity of graphene-based
[131] P.K. Namburu, D.P. Kulkarni, D. Misra, D.K. Das, Viscosity of copper oxide na- composites, Appl. Phys. Lett. 101 (2012).
noparticles dispersed in ethylene glycol and water mixture, Exp. Therm Fluid Sci. [162] R.P. Singh, H. Xu, S.C. Kaushik, D. Rakshit, A. Romagnoli, Charging performance
32 (2007) 397–402. evaluation of finned conical thermal storage system encapsulated with nano-en-
[132] R.S. Vajjha, D.K. Das, P.K. Namburu, Numerical study of fluid dynamic and heat hanced phase change material, Appl. Therm. Eng. 151 (2019) 176–190.
transfer performance of Al2O3 and CuO nanofluids in the flat tubes of a radiator, [163] R.P. Singh, H. Xu, S.C. Kaushik, D. Rakshit, A. Romagnoli, Effective utilization of
Int. J. Heat Fluid Flow 31 (2010) 613–621. natural convection via novel fin design & influence of enhanced viscosity due to
[133] J. Zhou, M. Hu, D. Jing, The synergistic effect between surfactant and nanoparticle carbon nano-particles in a solar cooling thermal storage system, Sol. Energy 183
on the viscosity of water-based fluids, Chem. Phys. Lett. 727 (2019) 1–5. (2019) 105–119.
[134] U. Nithiyanantham, L. González-Fernández, Y. Grosu, A. Zaki, J.M. Igartua, [164] K.W. Shah, Y. Lu, Morphology, large scale synthesis and building applications of
A. Faik, Shape effect of Al2O3 nanoparticles on the thermophysical properties and copper nanomaterials, Constr. Build. Mater. 180 (2018) 544–578.
viscosity of molten salt nanofluids for TES application at CSP plants, Appl. Therm. [165] A.A. Rabienataj Darzi, M. Jourabian, M. Farhadi, Melting and solidification of
Eng. 169 (2020) 114942. PCM enhanced by radial conductive fins and nanoparticles in cylindrical annulus,
[135] F. Wang, L. Han, Z. Zhang, X. Fang, J. Shi, W. Ma, Surfactant-free ionic liquid- Energy Convers. Manage. 118 (2016) 253–263.
based nanofluids with remarkable thermal conductivity enhancement at very low [166] S. Liu, Y. Li, Y. Zhang, Mathematical solutions and numerical models employed for
loading of graphene, Nanoscale Res. Lett. 7 (2012) 314. the investigations of PCMs‫ ׳‬phase transformations, Renew. Sustain. Energy Rev. 33
[136] J. Liu, F. Wang, L. Zhang, X. Fang, Z. Zhang, Thermodynamic properties and (2014) 659–674.
thermal stability of ionic liquid-based nanofluids containing graphene as advanced [167] M. Jourabian, M. Farhadi, A.A. Rabienataj Darzi, Outward melting of ice enhanced
heat transfer fluids for medium-to-high-temperature applications, Renew. Energy by Cu nanoparticles inside cylindrical horizontal annulus: Lattice Boltzmann ap-
63 (2014) 519–523. proach, Appl. Math. Model., 37 (2013) 8813-8825.
[137] J.C. Maxwell, A treatise on electricity and magnetism, Clarendon press, Oxford, [168] H.E. Patel, T. Sundararajan, T. Pradeep, A. Dasgupta, N. Dasgupta, S.K. Das, A
1873. micro-convection model for thermal conductivity of nanofluids, Pramana 65
[138] R.L. Hamilton, O.K. Crosser, Thermal Conductivity of Heterogeneous Two- (2005) 863–869.
Component Systems, Ind. Eng. Chem. Fundam. 1 (1962) 187–191. [169] M. Jourabian, M. Farhadi, K. Sedighi, On the expedited melting of phase change
[139] J. Koo, C. Kleinstreuer, A new thermal conductivity model for nanofluids, J. material (PCM) through dispersion of nanoparticles in the thermal storage unit,
Nanopart. Res. 6 (2005) 577–588. Comput. Math. Appl. 67 (2014) 1358–1372.
[140] S.K.N. Wakao, Heat and Mass Transfer in Packed Beds, Gordon and Breach Science [170] M. Jourabian, M. Farhadi, A. Rabienataj Darzi, Constrained ice melting around
Publications, New York, 1982. one cylinder in horizontal cavity accelerated using three heat transfer enhance-
[141] R. Elbahjaoui, H. El Qarnia, Transient behavior analysis of the melting of nano- ment techniques, Int. J. Therm. Sci. 125 (2018) 231–247.
particle-enhanced phase change material inside a rectangular latent heat storage [171] S.F. Hosseinizadeh, A.A.R. Darzi, F.L. Tan, Numerical investigations of un-
unit, Appl. Therm. Eng. 112 (2017) 720–738. constrained melting of nano-enhanced phase change material (NEPCM) inside a
[142] Y.M.F. El Hasadi, J.M. Khodadadi, Numerical simulation of the effect of the size of spherical container, Int. J. Therm. Sci. 51 (2012) 77–83.
suspensions on the solidification process of nanoparticle-enhanced phase change [172] N. Hajighafoori Boukani, A. Dadvand, A.J. Chamkha, Melting of a Nano-enhanced
materials, J. Heat Transfer 135 (2013). Phase Change Material (NePCM) in partially-filled horizontal elliptical capsules
[143] K. Khanafer, K. Vafai, M. Lightstone, Buoyancy-driven heat transfer enhancement with different aspect ratios, Int. J. Mech. Sci. 149 (2018) 164–177.
in a two-dimensional enclosure utilizing nanofluids, Int. J. Heat Mass Transf. 46 [173] M.T. Kohyani, B. Ghasemi, A. Raisi, S.M. Aminossadati, Melting of
(2003) 3639–3653. cyclohexane–Cu nano-phase change material (nano-PCM) in porous medium
[144] R.S. Vajjha, D.K. Das, Experimental determination of thermal conductivity of three under magnetic field, J. Taiwan Inst. Chem. Eng. 77 (2017) 142–151.
nanofluids and development of new correlations, Int. J. Heat Mass Transf. 52 [174] Z. Li, A. Shahsavar, A.A.A.A. Al-Rashed, P. Talebizadehsardari, Effect of porous
(2009) 4675–4682. medium and nanoparticles presences in a counter-current triple-tube composite
[145] C.H. Chon, K.D. Kihm, S.P. Lee, S.U.S. Choi, Empirical correlation finding the role porous/nano-PCM system, Appl. Therm. Eng. 167 (2020) 114777.
of temperature and particle size for nanofluid (Al2O3) thermal conductivity en- [175] M. Parsazadeh, X. Duan, Numerical study on the effects of fins and nanoparticles
hancement, Appl. Phys. Lett. 87 (2005). in a shell and tube phase change thermal energy storage unit, Appl. Energy 216
[146] C.-W. Nan, R. Birringer, D.R. Clarke, H. Gleiter, Effective thermal conductivity of (2018) 142–156.
particulate composites with interfacial thermal resistance, J. Appl. Phys. 81 (1997) [176] M. Ghalambaz, A. Doostani, A.J. Chamkha, M.A. Ismael, Melting of nanoparticles-
6692–6699. enhanced phase-change materials in an enclosure: Effect of hybrid nanoparticles,
[147] S. Harish, D. Orejon, Y. Takata, M. Kohno, Thermal conductivity enhancement of Int. J. Mech. Sci. 134 (2017) 85–97.
lauric acid phase change nanocomposite with graphene nanoplatelets, Appl. [177] A.J. Chamkha, A. Doostanidezfuli, E. Izadpanahi, M. Ghalambaz, Phase-change
Therm. Eng. 80 (2015) 205–211. heat transfer of single/hybrid nanoparticles-enhanced phase-change materials
[148] Z. Xu, C. Kleinstreuer, Concentration photovoltaic–thermal energy co-generation over a heated horizontal cylinder confined in a square cavity, Adv. Powder
system using nanofluids for cooling and heating, Energy Convers. Manage. 87 Technol. 28 (2017) 385–397.
(2014) 504–512. [178] M. Esapour, A. Hamzehnezhad, A.A. Rabienataj Darzi, M. Jourabian, Melting and
[149] M. Bechiri, K. Mansouri, Analytical study of heat generation effects on melting and solidification of PCM embedded in porous metal foam in horizontal multi-tube
solidification of nano-enhanced PCM inside a horizontal cylindrical enclosure, heat storage system, Energy Convers. Manage. 171 (2018) 398–410.
Appl. Therm. Eng. 104 (2016) 779–790. [179] R. Hossain, S. Mahmud, A. Dutta, I. Pop, Energy storage system based on nano-
[150] S. Shaikh, K. Lafdi, K. Hallinan, Carbon nanoadditives to enhance latent energy particle-enhanced phase change material inside porous medium, Int. J. Therm. Sci.
storage of phase change materials, J. Appl. Phys. 103 (2008) 94302. 91 (2015) 49–58.
[151] J.L. Zeng, Z. Cao, D.W. Yang, F. Xu, L.X. Sun, X.F. Zhang, L. Zhang, Effects of [180] J.M. Mahdi, E.C. Nsofor, Melting enhancement in triplex-tube latent heat energy
MWNTs on phase change enthalpy and thermal conductivity of a solid-liquid or- storage system using nanoparticles-metal foam combination, Appl. Energy 191
ganic PCM, J. Therm. Anal. Calorim. 95 (2009) 507–512. (2017) 22–34.
[152] J. Wang, H. Xie, Z. Xin, Thermal properties of paraffin based composites con- [181] Y. Pahamli, M.J. Hosseini, A.A. Ranjbar, R. Bahrampoury, Effect of nanoparticle
taining multi-walled carbon nanotubes, Thermochim Acta 488 (2009) 39–42. dispersion and inclination angle on melting of PCM in a shell and tube heat ex-
[153] N.H. Mohamed, F.S. Soliman, H. El Maghraby, Y.M. Moustfa, Thermal con- changer, J. Taiwan Inst. Chem. Eng. 81 (2017) 316–334.
ductivity enhancement of treated petroleum waxes, as phase change material, by α [182] S. Ebadi, S. Humaira Tasnim, A. Abbas Aliabadi, S. Mahmud, Geometry and na-
nano alumina: Energy storage, Renew. Sustain. Energy Rev. 70 (2017) 1052–1058. noparticle loading effects on the bio-based nano-PCM filled cylindrical thermal

28
T. Xiong, et al. Applied Thermal Engineering 178 (2020) 115492

energy storage system, Appl. Therm. Eng. 141 (2018) 724–740. J. Taiwan Inst. Chem. Eng. 67 (2016) 115–125.
[183] S. Ebadi, S.H. Tasnim, A.A. Aliabadi, S. Mahmud, Melting of nano-PCM inside a [202] M. Sheikholeslami, O. Mahian, Enhancement of PCM solidification using inorganic
cylindrical thermal energy storage system: Numerical study with experimental nanoparticles and an external magnetic field with application in energy storage
verification, Energy Convers. Manage. 166 (2018) 241–259. systems, J. Cleaner Prod. 215 (2019) 963–977.
[184] M. Parsazadeh, X. Duan, Numerical and statistical study on melting of nano- [203] M. Sheikholeslami Kandelousi, KKL correlation for simulation of nanofluid flow
particle enhanced phase change material in a shell-and-tube thermal energy sto- and heat transfer in a permeable channel, Phys. Lett. A 378 (2014) 3331–3339.
rage system, Appl. Therm. Eng. 111 (2017) 950–960. [204] J.M. Mahdi, E.C. Nsofor, Solidification of a PCM with nanoparticles in triplex-tube
[185] A.V. Arasu, A.S. Mujumdar, Numerical study on melting of paraffin wax with thermal energy storage system, Appl. Therm. Eng. 108 (2016) 596–604.
Al2O3 in a square enclosure, Int. Commun. Heat Mass Transfer 39 (2012) 8–16. [205] J.M. Mahdi, E.C. Nsofor, Solidification enhancement in a triplex-tube latent heat
[186] M. Arıcı, E. Tütüncü, M. Kan, H. Karabay, Melting of nanoparticle-enhanced energy storage system using nanoparticles-metal foam combination, Energy 126
paraffin wax in a rectangular enclosure with partially active walls, Int. J. Heat (2017) 501–512.
Mass Transf. 104 (2017) 7–17. [206] J.M. Mahdi, E.C. Nsofor, Solidification enhancement of PCM in a triplex-tube
[187] Z.-J. Zheng, Y. Xu, M.-J. Li, Eccentricity optimization of a horizontal shell-and- thermal energy storage system with nanoparticles and fins, Appl. Energy 211
tube latent-heat thermal energy storage unit based on melting and melting-soli- (2018) 975–986.
difying performance, Appl. Energy 220 (2018) 447–454. [207] H. Vatanparast, F. Shahabi, A. Bahramian, A. Javadi, R. Miller, The Role of
[188] A. Pizzolato, A. Sharma, K. Maute, A. Sciacovelli, V. Verda, Design of effective fins Electrostatic Repulsion on Increasing Surface Activity of Anionic Surfactants in the
for fast PCM melting and solidification in shell-and-tube latent heat thermal en- Presence of Hydrophilic Silica Nanoparticles, Sci. Rep. 8 (2018) 7251.
ergy storage through topology optimization, Appl. Energy 208 (2017) 210–227. [208] S. Sen, V. Govindarajan, C.J. Pelliccione, J. Wang, D.J. Miller, E.V. Timofeeva,
[189] S. Riahi, W.Y. Saman, F. Bruno, M. Belusko, N.H.S. Tay, Impact of periodic flow Surface Modification Approach to TiO2 Nanofluids with High Particle
reversal of heat transfer fluid on the melting and solidification processes in a latent Concentration, Low Viscosity, and Electrochemical Activity, ACS Appl. Mater.
heat shell and tube storage system, Appl. Energy 191 (2017) 276–286. Interfaces 7 (2015) 20538–20547.
[190] A. Sathishkumar, V. Kumaresan, R. Velraj, Solidification characteristics of water [209] H.W. Xian, N.A.C. Sidik, R. Saidur, Impact of different surfactants and ultra-
based graphene nanofluid PCM in a spherical capsule for cool thermal energy sonication time on the stability and thermophysical properties of hybrid nano-
storage applications, Int. J. Refrig 66 (2016) 73–83. fluids, Int. Commun. Heat Mass Transfer 110 (2020) 104389.
[191] L.-W. Fan, Z.-Q. Zhu, M.-J. Liu, A similarity solution to unidirectional solidifica- [210] B. Munkhbayar, M. Bat-Erdene, B. Ochirkhuyag, D. Sarangerel, B. Battsengel,
tion of nano-enhanced phase change materials (NePCM) considering the mushy H. Chung, H. Jeong, An experimental study of the planetary ball milling effect on
region effect, Int. J. Heat Mass Transf. 86 (2015) 478–481. dispersibility and thermal conductivity of MWCNTs-based aqueous nanofluids,
[192] R.P. Singh, S.C. Kaushik, D. Rakshit, Solidification behavior of binary eutectic Mater. Res. Bull. 47 (2012) 4187–4196.
phase change material in a vertical finned thermal storage system dispersed with [211] J. Wensel, B. Wright, D. Thomas, W. Douglas, B. Mannhalter, W. Cross, H. Hong, J.
graphene nano-plates, Energy Convers. Manage. 171 (2018) 825–838. Kellar, P. Smith, W. Roy, Enhanced thermal conductivity by aggregation in heat
[193] M. Alizadeh, K. Hosseinzadeh, M.H. Shahavi, D.D. Ganji, Solidification accelera- transfer nanofluids containing metal oxide nanoparticles and carbon nanotubes,
tion in a triplex-tube latent heat thermal energy storage system using V-shaped fin Appl. Phys. Lett., 92 (2008) 023110-023110-023113.
and nano-enhanced phase change material, Appl. Therm. Eng. 163 (2019) 114436. [212] Y. Hwang, J.K. Lee, C.H. Lee, Y.M. Jung, S.I. Cheong, C.G. Lee, B.C. Ku, S.P. Jang,
[194] J.M. Khodadadi, S.F. Hosseinizadeh, Nanoparticle-enhanced phase change mate- Stability and thermal conductivity characteristics of nanofluids, Thermochim Acta
rials (NEPCM) with great potential for improved thermal energy storage, Int. 455 (2007) 70–74.
Commun. Heat Mass Transfer 34 (2007) 534–543. [213] X.-J. Wang, X. Li, S. Yang, Influence of pH and SDBS on the Stability and Thermal
[195] M. Abdollahzadeh, M. Esmaeilpour, Enhancement of phase change material (PCM) Conductivity of Nanofluids, Energy Fuels 23 (2009) 2684–2689.
based latent heat storage system with nano fluid and wavy surface, Int. J. Heat [214] X. Li, D. Zhu, X. Wang, Evaluation on dispersion behavior of the aqueous copper
Mass Transf. 80 (2015) 376–385. nano-suspensions, J. Colloid Interface Sci. 310 (2007) 456–463.
[196] R. Elbahjaoui, H. El Qarnia, Thermal analysis of nanoparticle-enhanced phase [215] S. Salyan, S. Suresh, Study of thermo-physical properties and cycling stability of d-
change material solidification in a rectangular latent heat storage unit including Mannitol-copper oxide nanocomposites as phase change materials, J. Storage
natural convection, Energy Build. 153 (2017) 1–17. Mater. 15 (2018) 245–255.
[197] S. Tiari, S. Qiu, M. Mahdavi, Discharging process of a finned heat pipe–assisted [216] S. Madruga, G.S. Mischlich, Melting dynamics of a phase change material (PCM)
thermal energy storage system with high temperature phase change material, with dispersed metallic nanoparticles using transport coefficients from empirical
Energy Convers. Manage. 118 (2016) 426–437. and mean field models, Appl. Therm. Eng. 124 (2017) 1123–1133.
[198] S. Lohrasbi, S.Z. Miry, M. Gorji-Bandpy, D.D. Ganji, Performance enhancement of [217] N. Sezer, M.A. Atieh, M. Koç, A comprehensive review on synthesis, stability,
finned heat pipe assisted latent heat thermal energy storage system in the presence thermophysical properties, and characterization of nanofluids, Powder Technol.
of nano-enhanced H2O as phase change material, Int. J. Hydrogen Energy 42 344 (2019) 404–431.
(2017) 6526–6546. [218] Y.M.F. El Hasadi, Numerical simulation of the melting process of nanostructured
[199] S. Lohrasbi, M. Sheikholeslami, D.D. Ganji, Discharging process expedition of based colloidal suspensions phase change materials including the effect of the
NEPCM in fin-assisted Latent Heat Thermal Energy Storage System, J. Mol. Liq. transport of the particles, J. Mol. Liq. 287 (2019) 110886.
221 (2016) 833–841. [219] J. Lv, C. Hu, M. Bai, L. Li, L. Shi, D. Gao, Visualization of SiO2-water nanofluid
[200] S. Lohrasbi, M. Sheikholeslami, D.D. Ganji, Multi-objective RSM optimization of flow characteristics in backward-facing step using PIV, Exp. Therm Fluid Sci. 101
fin assisted latent heat thermal energy storage system based on solidification (2019) 151–159.
process of phase change Material in presence of copper nanoparticles, Appl. [220] S. Motahar, N. Nikkam, A.A. Alemrajabi, R. Khodabandeh, M.S. Toprak,
Therm. Eng. 118 (2017) 430–447. M. Muhammed, Experimental investigation on thermal and rheological properties
[201] M. Sheikholeslami, S. Lohrasbi, D.D. Ganji, Response surface method optimization of n-octadecane with dispersed TiO2 nanoparticles, Int. Commun. Heat Mass
of innovative fin structure for expediting discharging process in latent heat Transf. 59 (2014) 68–74.
thermal energy storage system containing nano-enhanced phase change material,

29

You might also like