You are on page 1of 21

Solar Energy 183 (2019) 632–652

Contents lists available at ScienceDirect

Solar Energy
journal homepage: www.elsevier.com/locate/solener

Thermal management of concentrator photovoltaic systems using new T


configurations of phase change material heat sinks
Ramy Rabiea,e, Mohamed Emamb, Shinichi Ookawaraa,c, Mahmoud Ahmeda,d,

a
Department of Energy Resources Engineering, Egypt-Japan University of Science and Technology (E-JUST), Alexandria 21934, Egypt
b
Department of Mechanical Engineering, Faculty of Engineering Shoubra, Benha University, Benha, Qalubiya 11629, Egypt
c
Department of Chemical Science and Engineering, Tokyo Institute of Technology, Tokyo 152-8552, Japan
d
Department of Mechanical Engineering, Faculty of Engineering, Assiut University, Assiut 71516, Egypt
e
Department of Mechanical Engineering, Faculty of Engineering, Suez Canal University, Ismailia 41522, Egypt

ARTICLE INFO ABSTRACT

Keywords: A new concentrator photovoltaic system integrated with phase change material heat sink is developed to achieve
Concentrator photovoltaic system low and uniform temperature distribution along the solar cell. The developed system includes a phase change
Thermal management material enclosure with different over height ratios, defined as the ratio between the over height length of the
Phase change material phase change material heat sink to the solar cell length. Subsequently, four different over height ratios of 0, 20,
Temperature uniformity
40, and 60 % along with three inclination angles of −45°, 0° and 45° at solar concentration ratios of 5 and 10 are
investigated. Thus, experimental and numerical investigations have been carried out to investigate the influence
of the over height ratio and inclination angle of the phase change material heat sink on the temperature dis-
tribution along the solar cell. To assess the performance of the suggested designs of heat sinks, a compendious
two-dimensional model of the concentrator photovoltaic cell layers combined with phase change material is
developed and numerically simulated. The numerical results are further validated with experimental measure-
ments. These measurements involve transient variation of liquid fraction and the evolution of the local tem-
perature during melting of RT35HC phase change material carried out at different inclination angles of −45°, 0°
and 45°. Results indicate that the influence of varying the over height ratio depends on the inclination angle of
PCM heat sink. It is found that at an inclination angle of −45° and a concentration ratio of 5, the peak cell
temperature reduces from 92 to 74 °C, and the temperature uniformity varies from 13.7 to 5.3 °C as the over
height ratio rises from zero to 60 %. However, at an inclination angle of 45° only, a slight reduction in the peak
cell temperature is observed along with a minor improvement of temperature uniformity. Results of the present
work can assist in the selection of appropriate phase change material -heat sink design for the required types of
solar concentrators.

1. Introduction overcome the excessive cost of regular PV modules setup via employing
inexpensive optical light concentrators like mirrors or lenses. The key
Since the oil crisis in the1970s, when photovoltaics (PV) were in- advantage of using concentrators is their ability to gather the incident
troduced as an innovative and clean technique to convert solar irra- solar irradiance from a wide area and concentrate that energy onto
diance directly to an electrical current, extensive research has been small area. This intensely lowers the utilization of high-priced semi-
directed to this field. This has resulted in a substantial cost reduction of conductors materials in which the system cost is also reduced. In ad-
the PV per wattage from 200 $ in 1970 s to less than 0.5 $ in 2017 dition, CPV systems offer more advantages such as large electricity gain
(Horowitz et al., 2017) and also boosted their electrical efficiency to compared to conventional PV systems, possibility of recycling the em-
higher than 20 % compared to 2.0 % in 1950s (Hedayatizadeh et al., ployed materials and reducing the use of toxic components related to
2013). However, the initial cost of PV systems is still high and remains a the manufacturing and disposal of PV cells.
significant roadblock to wide-spread utilization and impede competi- Concentrator photovoltaic systems can be subjected to both low
tion with regular power generated from fossil fuels (Green, 2013). concentrated solar irradiance (< 100 suns) and high concentrated solar
Development of concentrator photovoltaics (CPV) systems helped irradiance (> 100 suns) (Sarwar et al., 2014). Moreover, for crystalline


Corresponding author at: Department of Mechanical Engineering, Assiut University, Assiut 71516, Egypt.
E-mail addresses: aminism@aun.edu.eg, mahmoud.ahmed@ejust.edu.eg (M. Ahmed).

https://doi.org/10.1016/j.solener.2019.03.061
Received 26 November 2018; Received in revised form 9 March 2019; Accepted 15 March 2019
Available online 21 March 2019
0038-092X/ © 2019 International Solar Energy Society. Published by Elsevier Ltd. All rights reserved.
R. Rabie, et al. Solar Energy 183 (2019) 632–652

Nomenclature β1 thermal expansion coefficient [1/K]


τj transmissivity of layers above the layer i
A area [m2] η efficiency
Amush mush zone constant [kg/m3 s] emissivity
a over height length (m) ω the uncertainty
BIPV building integrated photovoltaic μ viscosity [Pa s]
Cp specific heat [J/kg K] ρ density [kg/m3]
CR concentration ratio σ Stephan–Boltzmann constant 5.67 × 10−8 [W/m2 K4]
CPV concentrator photovoltaic δ thickness [m]
EVA ethylene vinyl acetate τ transmissivity
g gravity acceleration [m/s2] Θ inclination angle
G(t) concentrated solar irradiance [W/m2] Λ over height ratio
h heat transfer coefficient [W/m2 K]
href reference enthalpy [J/kg] Subscripts
H total enthalpy [J/kg]
K thermal conductivity [W/m K] amb ambient
L latent heat [J/kg] al aluminum
P pressure (Pa) conv convection
PV photovoltaic E EVA layer
PCM phase change material g glass
qi internal heat generation per unit volume [W/m3] l liquid
Sx source term in x direction m melting
Sy source term in y direction max maximum
t time [min] min minimum
T temperature [°C] rad radiation
TCE thermal conductivity enhancer ref reference
u, v velocity in x and y-direction respectively [m/s] s solid
V velocity [m/s] and volume [m3] sc solar cell
sky sky
Greek symbols T tedlar
th thermal
αi absorptivity of PV layer i wind wind
αl thermal diffusivity of liquid [m2/s] x x-direction
αs thermal diffusivity of solid [m2/s] y y-direction

silicon solar cells that are currently used in CPV, only 20 % of the cost is needed compared to active cooling methods.
concentrated solar irradiance is converted into electricity while the Numerous investigations have been conducted to evaluate the in-
remainder is transformed into heat energy (Sharma et al., 2017), re- corporation of PCM with the conventional flat PV systems under normal
sulting in excessive increase in the CPV cell temperature. This situation operating conditions. Hasan et al. (2017) studied the integration of
reduces the electrical power production and may lead to permanent paraffin PCMs of melting temperature (38–43) °C with flat PV panel in
damage of the cell especially at high concentration ratios. Conse- dry hot weather of United Arab Emirates through the entire year. They
quently, effective thermal management of concentrator photovoltaic found that the PCM helped to enhance the PV annual electricity output
systems is a mandatory. Accordingly, many researchers have in- by about 5.9 % with a maximum decrease in the temperature of the cell
vestigated and introduced various cooling techniques for CPV systems up to 13 °C. Moreover, Hasan et al. (2010) experimentally examined the
and they recommend that a proper cooling method should maintain the utilization of five different PCMs with melting temperatures of
silicon cell layer at a reduced temperature to boost the electrical per- 25 ± 4 °C and latent heat ranging from 140 to 213 kJ/kg in cooling of
formance of the system, protect the system from fast degradation due to building integrated photovoltaics (BIPV) systems. Results indicated that
excessive thermal stresses and ensure temperature uniformity along the PCM-based cooling of PV systems depend on thermal conductivity, la-
CPV cell surface (Radwan et al., 2019; Radwan et al., 2018; Radwan tent heat and thermal mass of the employed PCM. Stropnik and Stritih
and Ahmed, 2017). In addition, it should exhibit low operating and (2016) numerically and experimentally studied the effect of in-
maintenance costs to make the CPV remain competitive with regular corporating PCM in flat PV on the output power and the electrical ef-
flat PV (Browne et al., 2015). Another important fact is the usability of ficiency. They reported that the solar cell temperature decreased by
the extracted thermal energy from the system. Several CPV cooling maximum of 35.6 °C using PCM. In addition, their results indicated that
techniques are available such as passive cooling techniques including about a 7.3 % increase in the annual electricity production was sus-
phase change material (PCM) and natural air circulation, active cooling tained under the climate conditions of Ljubljana, Slovenia. Khanna
techniques including hydraulic cooling (Dong et al., 2018), spray et al. (2018) examined the performance of PV integrated with PCM
cooling (Nizetic et al., 2016) forced air convection and thermoelectric under different climatic conditions. They reported a 9.7 % improve-
(Peltier effect) cooling (Royne et al., 2005). The focus of the current ment in electricity production for a climate with low variation in am-
research is on passive cooling method by integrating PCM with CPV bient temperature and a 6.7 % for a climate with large variation in
systems. Phase change materials can absorb and release a huge quantity ambient temperatures.
of latent thermal energy at almost constant temperature during its The main obstacle of utilizing PCMs as a heat sink is their typically
phase transition from solid to liquid or vice versa (Ahmed and Radwan, low thermal conductivities that significantly limit their cooling per-
2017; Radwan et al., 2017). Furthermore, PCM-based cooling systems formance. Additionally, due to the melting behavior of PCMs, the hot
are noiseless, do not need power to operate and minimum maintenance liquid phase is amassed at the top portion of the PCM container

633
R. Rabie, et al. Solar Energy 183 (2019) 632–652

resulting in the formation of stratified liquid layers. This situation leads CPV at an angle of 45° achieved the lowest solar cell temperature and
to nonuniformity of the temperature variation along the CPV cell, and consequently the best electrical performance. However, an opposite
eventually creates hot spots in the upper part of the solar cells. trend was observed at an inclination angle of −45° which showed the
Therefore, several techniques had been employed to improve the heat highest temperature and the lowest electrical performance of the CPV-
transfer process in PCMs and in turn reduce the temperature non- PCM system. Additional investigation with respect to inventing novel
uniformity of the CPV-PCM system. Some researchers utilized thermal and efficient designs is essential to advance such integrated systems.
conductivity enhancers (TCEs) such as metallic fins, nanoparticles, and Based on the literature survey, it is concluded that a little research
metal foams in boosting the performance of PCM-heat sinks used for works have been conducted to investigate the formation of stratified
thermal energy storage and thermal management of CPV systems. liquid layers associated with the melting behaviour of PCMs. This es-
Kamkari and Shokouhmand (2014) experimentally investigated the sentially causes a temperature nonuniformity along the solar cell and
melting behavior of lauric acid (PCM) in a rectangular cavity with and generates potential hot spots in the upper part of the solar cells.
without partial metal fins. Their measurements revealed that increasing Therefore, in the current work, a novel design of CPV-PCM system is
the number of partial fins reduced the time required to reach the suggested to attain low and uniform temperature distribution along the
complete melting phase and increased the total heat transfer rate inside solar cell. The developed system includes a PCM enclosure with dif-
the PCM domain. Moreover, the thermally stratified part reduced as the ferent over height ratios (Λ). Thus, different over height ratios of 0, 20,
number of fins was raised in which the temperature uniformity was 40, and 60 % with various inclination angles of −45°, 0° and 45° that
enhanced. Modifying the PCM heat sink design is a promising way for are appropriate for solar concentrators at solar concentration ratios
improving the thermal management of CPV systems. Emam and Ahmed (CRs) of 5 and 10 are investigated. To assess the performance of the
(2018a) proposed a new integrated CPV-PCM system to boost the new suggested design of heat sink, a comprehensive two-dimensional
thermal conductivity via using various configurations of the PCM heat (2-D) mathematical model for the CPV cell layers combined with PCM
sink and improve their cooling performance. They indicated that the was developed and numerically simulated.
integrated CPV-PCM system with a five-parallel cavity configuration The model can track the liquid-solid interface evolution and predict
heat sink attained the best CPV performance because of the enhanced the temperature field of the PCM and the solar cell. Based on the ob-
heat transfer inside the PCM domain. However, the used fins act like tained results, the over height ratio and inclination angle that showed
obsticles to the natural flow of the liquid phase and possibly reduce its the best CPV performance are identified. To validate the developed
natural convective heat transfer. In addition, fins occupy part of the model at different inclination angles of CPV-PCM system, the effects of
volume rather than the PCM resulting in the reduction of thermal the inclination angle on CPV-PCM systems performance are experi-
regulation period (Huang et al., 2011). Dhaidan et al. (2013) carried mentally investigated to identify the best orientation of such integrated
out an experimental and numerical study to assess the effect of adding systems. Consequently, the thermal behaviour of the melting of
cupric oxide (CuO) nanoparticles with different concentrations to n- RT35HC-PCM at different inclination angles (Θ) of −45°, 0° and 45°
octadecane PCM on its thermal behavior during melting. It was con- and a heat flux value of 3000 W/m2 are experimentally investigated.
cluded that the nanoparticles loading showed an encouraging trend in The fabricated heat sink is attached to a silicon rubber heater to mimic
boosting the thermal conductivity of the PCM-nanoparticle composite the heat dissipation from CPV cells in CPV-PCM systems.
and rising the rate of heat transfer which resulted in reduction of the
charging time. Zarma et al. (2019) studied the influence of different 2. Methodology
nanoparticles (Al2O3, CuO and SiO2) on the performance of a con-
centrator photovoltaic system. Their results concluded that PCM’s The current study starts with developing a comprehensive 2-D
thermal conductivity improved with the addition of Al2O3 which en- model to assess the thermal behavior of a CPV-PCM system at different
hanced the heat transfer process, rate of melting, and consequently over height ratios of 0, 20, 40, and 60 % with various inclination angles
decreases the solar cell temperature. Another experimental and nu- of −45°, 0° and 45° at solar concentration ratios (CRs) of 5 and 10. This
merical study to assess thermal regulation of PV systems using PCM model is able to elucidate the internal thermo-fluid phenomena during
infused with graphite in an externally-finned heat sink was accom- melting of the PCMs that were not captured in the experiments. To
plished by (Atkin and Farid, 2015). Their results indicated that the validate the developed model along with further exploration of the
thermal conductivity of PCM was improved from 0.25 W/m K to effect of wall tilting angle on the temperature variation of a CPV-PCM
16.6 W/m K due to the infusion of PCM into graphite along with the system, a fabricated PCM heat sink was attached to a silicon rubber
attached fins. This increased the PV panel overall efficiency by about heater to emulate the heat dissipation from CPV cells. The system was
13%. Zhu et al., (2018) examined the effect of filling height ratio and experimentally investigated at inclination angles of −45°, 0° and 45°
pore size on the performance of PCM-heat sink filled with a copper during melting of RT35HC-PCM at a heat flux value of 3000 W/m2.
foam used for electronic devices cooling. They reported that a tem-
perature reduction of up to 40 % was achieved when two thirds of PCM 2.1. Theoretical analysis
enclosure was filled with the copper foam with pore size of 15 pores per
inch (ppi) under heating power of 80 W. However, utilizing metal fins In this section, a comprehensive 2-D model for of CPV-PCM system
or foams to enhance the heat transfer process inside the PCM domain was developed to determine the transient average and local tempera-
raised the system weight and cost. ture variation of a solar cell at different operating conditions. Then,
On the other hand, other researchers enhanced the cooling perfor- numerical simulations were performed to determine the effects of
mance of PCMs via boosting natural convection currents during the adding a mass of PCM over the solar cell height on cooling of the CPV
solid-liquid phase change process. Rabie et al. (2017) studied the systems. This is accomplished through changing the PCM cavity over
thermal behaviour of a CPV system integrated with a modified paral- height ratio, Λ while keeping the PCM volume and the solar cell height
lelogram shaped PCM heat sink with those of a regular rectangular constant. The over height ratio is calculated using Eq. (1) and expressed
shape. Their numerical results showed that the new design attained a as the ratio between the over height length of the PCM heat sink (a)
temperature reduction of up to 9 °C when compared to the system with (which is defined as the distance between the end of the solar cell and
regular rectangular shape because of the enhanced heat transfer process the end of PCM container) to the solar cell length (h). The model con-
during melting of PCM. Emam et al. (2017a) numerically inspected the siders these modifications in three different inclination angles of −45°
thermal behavior and electrical performance of a CPV system integrated (heated side on top), 0° (vertical) and 45° (heated side on bottom) to
with PCM at different inclination angles (−45° to 90°) and PCM simulate possible orientations required for different types of solar
thickness (5 cm and 20 cm). They found that incorporating PCM with concentrators such as dish concentrator, compound parabolic trough

634
R. Rabie, et al. Solar Energy 183 (2019) 632–652

and Fresnel lens. Four different over height ratios of 0, 20, 40, and 60 % proposed CPV-PCM system are written as follows:
are investigated at each inclination angle under two distinct values of
CR of 5 and 10.
2.1.1. Photovoltaic layers and aluminum sheets
a Energy conservation in solid parts was written as (Emam and
=
h (1) Ahmed, 2018b):

A neat sketch of the CPV-PCM configuration considered in the present Ti (x , y ) 2T (x , y ) 2T (x , y)


i Cp, i = ki ( i 2 + i
) + qi where i = 1, 2, ..,n
work with geometrical variables and boundary conditions is depicted in t x y2
Fig. 1. In this study, a generic polycrystalline cell as defined by Sandia (2)
national laboratory is investigated. The cell contains a silicon layer
encapsulated between two ethylene vinyl acetate (EVA) sheets, a glass where Cp,i, ρi, Ti(x, y), ki and qi are the specific heat capacity (J/kg K),
cover at the top and a tedlar back sheet layer as shown in Fig. 1. density (kg/m3), temperature (°C), thermal conductivity (W/m °C) and
Moreover, it is directly connected to the PCM heat sink front surface internal heat generation (W/m3) for layer i, respectively.
that is fabricated from 3 mm thick aluminum sheets. These aluminum In each layer of the CPV cell, the absorbed solar irradiance can be
walls attain an even temperature distribution over the cell surface and estimated using Eq. (3) and substituted in Eq. (2) as an internal heat
offer the capability to shield the PCM. generation per unit volume of the layer i, qi.
In the present study, the concentrated solar irradiance on the CPV-
(1 e )·G(t)· i· j ·Ai
PCM system is assumed to be equally distributed. This assumption is qi =
Vi (3)
considered in many previous literatures, for instance (Kant et al.,
2016a, 2016b; Radwan et al., 2016b; Zhou et al., 2015). However, in where: ηe represents the silicon cell layer electrical efficiency and equals
the real case, the intensity distribution is expected to be slightly peaked zero for other layers; G(t) is the incident solar irradiance (W/m2) and τj
in the centre due to the optical design of concentrators. In authors’ represents the transmissivity of layers above layer i; Ai, αi and Vi are the
opinions, this situation has a minimal consequence on the output results area (m2), absorptivity and volume (m3) of layer i, respectively. Further
particularly in the case of the cooled CPV systems. Furthermore, in the explanations of the calculation of qi for each CPV layer are described in
case of the CPV-PCM system, an aluminum wall with high thermal (Emam and Ahmed, 2018a).
diffusivity was integrated with the rear side of the CPV cell to achieve
almost uniform flux distribution over the front surface of the system
and provided a high rate of heat transfer so that heat was absorbed 2.1.2. Phase change material
quickly from the CPV cell. This design will aid in minimizing the effect The enthalpy-porosity formulation is employed to model the PCM
of existing peak of solar intensity. melting process where the liquid fraction term (λ) is used to follow the
The developed model combines a heat conduction model for CPV existence of liquid or solid phases. Appropriate source terms (Sx and Sy)
layers and a transient thermal fluid model that includes the phase- are substituted in momentum equations to model the velocity changes
change phenomenon using the enthalpy porosity formulation. In addi- from the stationary solid-phase region to the liquid-phase region. The
tion, the model can track the development of the transient liquid-solid Boussinesq approximation is applied to include thermal buoyancy in
interface, the temperature field of the PCM during the melting process the liquid PCM. The 2-D transient analysis of phase change process is
and the temperature distribution of the CPV cell at different working written as follows (Brent et al., 1988; Pal and Joshi, 2001):
conditions. The governing equations for each component of the Continuity equation:

Fig. 1. A neat sketch of the studied CPV-PCM system.

635
R. Rabie, et al. Solar Energy 183 (2019) 632–652

Table 1 numerically (Dhaidan et al., 2013).


Time step and mesh independence test.
Tm = Tsolidus Tliquidus (14)
Grid no. Time step Liquid fraction Average solar cell temperature (K)
Moreover, for the phase change materials utilized in the simulation and
49,408 0.2 0.484 356.51 experimental work (RT35 HC and RT44HC), the thermal conductivity is
94,062 0.2 0.518 350.89 constant for both the solid and the liquid phases according to the da-
165,912 0.2 0.532 348.2
165,912 0.1 0.532 348.6
tasheet of the manufacturer (“Rubitherm GmbH,” 2014). However, for
248,382 0.2 0.534 348.7 other PCM used in the validation with (Kamkari et al., 2014) results
(Lauric acid), the PCM thermal conductivity varies only from 0.16 W/m
k to 0.14 W/m K with the phase transition from solid to liquid phase. In
u
+
v
=0 that case, Eq. (13) is used to simulate this slight change.
x y (4) Finally, the uniformity index of solar cell, ΔT which is defined as the
difference between the local maximum and minimum temperature of
Momentum equations:
the solar cell is calculated using the following equation:
u u u 1 P µ 2u 2u
+u +v = + + + g (T Tref ) cos T = Tsc,max Tsc ,min (15)
t x y x x2 y2
+ Sx (5)
2.1.3. Initial and boundary conditions
v v v 1 P µ 2v 2v Initially, the system contains a solid PCM maintained at a tem-
+u +v = + + + g (T Tref ) sin + Sy perature (Tini) lower than the melting temperature (Tm) of the employed
t x y y x2 y2
PCMs, and equals to 25 °C. Furthermore, no-slip boundary conditions
(6) are considered at the solid-fluid interfaces. However, for all existing
1 (1 )2 solid-solid interfaces, a thermally coupled boundary condition is ap-
Sx = · 3 · Amush · u
( + ) (7) plied. The heat gained due to the absorbed solar irradiance of each
layer in the CPV cell is calculated using Eq. (3) and applied to the heat
1 (1 )2 transfer equation as internal heat generation, qi. In addition, adiabatic
Sy = · 3 · Amush · v
( + ) (8) boundary condition is applied on the upper and lower ends of the CPV-
PCM system, as shown in Fig. 1. A thermal boundary condition for the
where μ and ρ are the dynamic viscosity (Pa s) and the density (kg/m3)
front surface of the PV cell is combined convective and radiative losses.
of the liquid PCM, respectively; g, β and Θ are the gravitational accel-
The convective heat transfer coefficient, ambient temperature, surface
eration (m/s2), the thermal expansion coefficient (1/K) and the in-
external emissivity, and the external radiation temperature are accu-
clination angle; λ and Amush are the liquid fraction and the mushy zone
rately specified. The exterior back boundary is subjected to a convective
constant, respectively. More details in regard to Amush are defined
heat loss for the CPV-PCM system. In detail, the applied boundary
elsewhere (Emam and Ahmed, 2018a). During simulations Amush is as-
conditions for the front and exterior back surfaces can be expressed as
sumed to be 106. Finally, γ is a small number of 10−3 to circumvent
follows (Radwan and Ahmed, 2018; Soliman and Hassan, 2019; Xu and
division by zero.
Kleinstreuer, 2014):
Equation of energy for the liquid phase is written as follows:
On the front surface of the CPV cell:
H H H H H At: x = 0 and (0 ≤ y ≤ h)
+u +v = l + l
t x y x x y y (9) T
kg = hconv, g amb (Tamb Tg |x ) + hrad, g sky (Tsky Tg |x )
Equation of energy for the solid phase could be formulated as follows: x x (16)

H H H On the exterior back surface of the CPV-PCM system:


= s + s At: x = (δg + 2δE + δsc + δT + 2 δal + b) and 0 ≤ y ≤ (h + a)
t x x y y (10)
The enthalpy of the PCM, H (J/kg) is calculated as: T
kal = hconv, al amb (Tamb Tal |x )
x x (17)
T
H = href + cp dT + L
Tref (11) where δg, δE, δsc, δT and δal are the thickness of glass, EVA, silicon cell,
Tedlar and aluminum layers. Tamb is the ambient temperature, and it is
where href is the reference enthalpy at reference temperature Tref; L is
assumed to be 25 °C (298 K).
the latent heat of PCM (J/kg).
The convective heat transfer coefficient from the glass cover to the
The liquid fraction, λ, is defined as:
atmosphere (hconv,g-amb) due to the wind effect was estimated by the
H
=0 T Tm following correlation (Agrawal and Tiwari, 2011; Radwan et al., 2016a;
L
Rejeb et al., 2016; Soliman and Hassan, 2019):
H T - Tsolidus
= = T Tm < T < Tm + Tm
L liquidus - Tsolidus hconv, g amb = 5.82 + 4.07 Vwind (18)
H
=1 T Tm + Tm (12)
L where Vwind is the wind speed, which is considered to be 1 m/s through
The thermal conductivity is taken as the following: all simulations.
This correlation was developed based on the fundamentals of heat
ks T Tm transfer theory and wind tunnel measurements. In these tests, a heated
kpcm = (ks + kl )
Tm < T < Tm + Tm plate was placed outdoors, and the forced convective heat transfer
2
coefficient was correlated with the prevailing wind speed. The wind
kl T Tm + Tm (13)
speed was measured at 1.0 m higher than the heated plate. All constants
where Tm is the melting temperature of the PCM, and ΔTm is the phase in this equation have dimensions and the equation is valid for a wind
transition range which is defined as the difference between the liquidus speed up to 5 m/s.
and solidus temperatures as shown in Eq. (14). This modification is The radiative heat transfer coefficient between the glass cover and
necessary since melting at an exact temperature cannot be modelled the sky (hrad,g-sky) is obtained using the following equation, based on the

636
R. Rabie, et al. Solar Energy 183 (2019) 632–652

assumption that the sky is a black body with a temperature Tsky (Emam reach the predefined accuracy within a reasonable time. Thus, four
et al., 2017b; Rejeb et al., 2016, 2015): distinct grid sizes were tested with consideration to average solar cell
temperature and the liquid fraction as shown in Table 1 and a mesh size
hrad, g = (Tg2 + Tsky
2
)(Tg + Tsky ) Tg and Tsky are in (K) (19)
sky g
of 165,912 is found to be the proper selection as increasing the grid size
where ɛg, σ are the emissivity of the glass cover and the Stefan-Boltz- to 248,382 elements showed a very small difference in the average
mann constant, respectively. The sky temperature Tsky is estimated temperature as well as in the liquid fraction. A time step of 0.2 sec was
using the ambient temperature Tamb as follows (Rabie et al., 2017): chosen after performing time step independence test of 0.2 and 0.1 s.
The tolerances were 10−5 for continuity and 10−6 for all the other
Tsky = Tamb 6 (20) equations.

The values of CPV cell characteristics and design parameters used in the
current work are described elsewhere (Emam and Ahmed, 2018a). 2.2. Experimental analysis

2.1.4. Numerical method 2.2.1. Experimental setup and procedures


The developed model combines a heat conduction model for CPV The setup of the main components used in the current study is
layers and a transient thermal fluid model that includes the phase schematically shown in Fig. 2. A rectangular cavity with internal di-
change phenomenon using the enthalpy porosity formulation. The mension of 50 mm in width, 120 mm in height, and 100 mm in depth
model is numerically simulated using ANSYS FLUENT 17.2 as described was employed as the PCM container. A silicon rubber heater with a
in detail (Emam and Ahmed, 2018b). The model characteristics are maximum power density of 0.5 W/cm2 was used as a constant heat flux
carefully studied based on a mesh and time step independence test to source to imitate the heat dissipation from CPV cells in CPV-PCM

Fig. 2. (a) A neat sketch of the experimental setup, (b) Real photos of the PCM enclosure and (c) A three-dimensional schematic view of the enclosure assembly.

637
R. Rabie, et al. Solar Energy 183 (2019) 632–652

Table 2
Thermo-physical properties for the employed PCMs, aluminum and plexiglass.
Thermo-physical properties RT35HC, for experiments and numerical RT44HC for experimental Aluminum Plexiglass
simulations validation

Melting temperature (range) (°C) 34–36 (main peak: 35) 41–44 (main peak: 43) N/A N/A
Heat of fusion (kJ/kg) 240 250 N/A N/A
Thermal conductivity
Solid (W/m °C) 0.2 0.2 211 0.19
Liquid (W/m °C) 0.2 0.2 N/A N/A
Density
Solid (kg/m3) 880 (at 25 °C) 800 (at 25 °C) 2675 1180
Liquid (kg/m3) 770 (at 40 °C) 700 (at 80 °C) N/A N/A
Specific heat capacity
Solid (kJ/kg K) 2 2 0.903 1.464
Liquid (kJ/kg K) 2 2 N/A N/A
Volumetric expansion, ΔV (1/kg) % 12 12.5 N/A N/A
Sub-cooling (°C) Negligible Negligible N/A N/A
Ref. (“Rubitherm GmbH,” 2014) (“Rubitherm GmbH,” 2014) (Kamkari and Groulx, (Kamkari and Groulx,
2018) 2018)

N/A: Not applicable.

Fig. 4. Measured and numerical results of liquid fraction variation during


melting of RT44HC PCM at a heat flux of 3750 W/m2.

high conductive aluminum plate (k = 218 W/m K). The remaining five
walls of the PCM containment were fabricated from 20 mm thick acrylic
clear Plexiglass clear plates. This is to facilitate visualization and pho-
tography of the PCM phase transition process and reduce heat loss to
the environment by their low thermal conductivity (k = 0.19 W/m K).
Acrylic plates were assembled together using M4 stainless steel bolts,
while silicon rubber thin sheets (0.5 mm thickness) were placed be-
tween the plates to prevent leakage. For further insulation, a 30 mm
thick foaming Urethane sheets (k = 0.026 W/m K) were used on all
sides of the enclosure with the aim of minimizing heat losses to the
environment.
Four calibrated K-type micro-thermocouples, (AS ONE Corp., Model
Fig. 3. The evolution of solid-liquid interface and photographs of RT44HC PCM
No. T1-SP-K, (“AS ONE,” 2018)), of 0.127 mm bare wire diameter were
melting at different times (heat flux = 3750 W/m2).
utilized for measurement of temperatures. Three thermocouples were
attached to the front aluminum plate to measure the transient local
systems. Electrical power was supplied to the heater via a volt slider, temperature variation of the system front wall as shown in Fig. 2. The
(AS ONE Corp., Model RSA-5, (“AS ONE,” 2018)), to control the remaining thermocouple was located outside the test cell for recording
amount of heat flux, while the wall heat flux was detected using a heat the ambient temperature. A multichannel temperature recorder was
flux sensor (Captec Enterprise, HF-10, (“Captec - heat flux sensor,” employed to record the temperatures at every one second. During the
2018)). To mount the heater and attain uniform heat flux on the system experiments, the melting of a commercially available PCM-RT35HC
surface, the front wall of the enclosure was formed from a 10 mm. thick (organic paraffin wax based) with a melting temperature of 35 °C was

638
R. Rabie, et al. Solar Energy 183 (2019) 632–652

Fig. 5. Measured and predicted results of the average front wall during melting Fig. 7. Comparison of temperature at point A between experimental and pre-
of RT44HC PCM at a heat flux of 3750 W/m2. dicted results.

the solid-liquid volume change during the melting process and to fa-
cilitate thermocouples fixation as shown in Fig. 2. The experiments
were initiated by supplying power to the electric heater via a volt slider
to adjust the heat flux value. Concurrently, the temperatures were
logged using the multichannel temperature recorder. Moreover, it is
worth mentioning that part of the insulation was periodically removed
for a time span of 8–9 s every 10 min from one side of the container and
photographs taken. This was to visualize and capture the liquid-solid
interface evolution and facilitate the calculation of liquid fraction of
PCM during melting process. The specifications and accuracy of the
employed instruments as reported in their datasheets are presented in
Appendix A (Table A.1).

2.3. Uncertainty analysis

The uncertainty in a derived parameter, X, due to the uncertainty in


the individual measured variables, x1, x2… xn, is referred to uncertainty
propagation. The total uncertainty in derived parameter X can be cal-
culated using the following expression (Radwan et al., 2018).
2 2 2
X 2 X 2 X 2
x = x1 + x2 + .......+ xn
x1 x2 xn (21)
Fig. 6. Measured and predicted results of liquid fraction.
where ωx is the uncertainty of the variable X; ωxn is the uncertainty in
observed at different tilting angles of −45°, 0° and 45°. However, parameter xn, and ∂X/∂x1 is the partial derivative of X with respect to
RT44HC and RT35HC-PCMs were used to validate the developed model x1. The estimated uncertainty in temperature is ± 0.5 °C for the ther-
with the current measurements and for numerical simulations, respec- mocouples. The uncertainty in length, width and area are ± 0.1
tively. The data sheet provided by the manufacturer (“Rubitherm mm, ± 0.1 mm and ± 4.1 mm2. The maximum value of uncertainty in
GmbH,” 2014) assures that the employed PCMs possesses a remarkable input heat flux is 4.5 %.
latent heat capacity in narrow temperature ranges, a stable perfor-
mance throughout repeated phase change cycles and a negligible super- 2.4. Model validations
cooling effect. Moreover, they are chemically inert, friendly to the en-
vironment, non-toxic, easy to handle, relatively low cost (about US$ Experimental measurements were carried out to test and validate
7.5/kg), and have an unlimited lifetime. Table 2 lists the thermo-phy- the developed model by comparing the numerically estimated results
sical properties of the utilized PCM and other components. with the current measurments and other available data for which the
Initially, the solid PCM was melted completely in a hot water bath same system geometry, boundary conditions, and material properties
and then well stirred before being poured gradually into the container were used. Initially, comparisons between the predicted results and the
in a layer by layer method to reduce the formation of air pockets inside current experimental measurments using RT44HC PCM are presented in
the solid PCM. A longitudinal hole (4 mm diameter) was drilled through Figs. 3–5. The exterior surface of the front aluminum wall was kept at a
the upper wall of the enclosure to empty the excess liquid-PCM due to constant heat flux of 3750 W/m2, while the remaining walls of the
Plexiglass were insulated. Fig. 3 shows a comparison between the

639
R. Rabie, et al. Solar Energy 183 (2019) 632–652

Fig. 8. Sequential photographs of the solid-liquid phase of RT35HC PCM for different orientation angles at a heat flux of 3000 W/m2 and Tamb = 25 °C.

estimated time of the liquid-solid interface evolution and experimental insignificance of the sharp corners. Furtheremore, the current estimated
images during melting of RT44HC PCM. Moreover, the predicted liquid results were compared with the corresponding experimental results of
fraction variation versus time and the evolution of the average front (Kamkari and Shokouhmand, 2014) as shown in Figs. 6 and 7. The
wall temperature during melting of RT44HC PCM at a heat flux value of comparison includes the current predicted melting fraction variation
3750 W/m2 are compared with the measurements as shown in Figs. 4, over time during melting of lauric acid (PCM) and the local temperature
and 5, respectively. These comparisons show that the 2-D numerical at point (A). It is illustrated that the predicted results using the present
simulations of PCM melting correctly predicted both the liquid fraction model are in good agreement with the obtained measurements and
and the liquid-solid interface evolution along with the transient varia- those other measurements.
tion of front wall temperature. However, the sharp corners (at 60 min,
70 min, 80 min) do commonly appear while simulating the melting of
phase change materials using computational fluid dynamics commer- 3. Results and discussion
cial packages as reported in previous studies (Zeng et al., 2017), and
(Kamkari and Amlashi, 2017). These differences have insignificant ef- This section comprises two main subsections. The first one reveals
fect on the obtained results. In addition, comparisons between mea- the effects of the inclination angle on CPV-PCM systems performance
surements and predicted values of the liquid fraction and the transient based on experiments to identify the best orientation of such integrated
variation of front wall temperature in Figs. 4 and 5 confirm systems. The melting of RT35HC PCM at various orientation angles (Θ)
of −45°, 0° and 45° and a heat flux value of 3000 W/m2 are

640
R. Rabie, et al. Solar Energy 183 (2019) 632–652

Fig. 9. The solid-liquid interface evolution of RT35HC PCM with different orientation angles at a heat flux value of 3000 W/m2 and Tamb = 25 °C.

Fig. 10. Transient liquid fraction during melting of RT35HC PCM at different Fig. 11. Transient average hot wall temperature during melting of RT35HC at
orientation angles at a heat flux value of 3000 W/m2 and Tamb = 25 °C. different orientation angles at a heat flux value of 3000 W/m2 and
Tamb = 25 °C.
demonstrated. The second subsection demonstrates the effects of
changing the PCM enclosure over height ratio (Λ) on the thermal Therefore, the cooling technique of inclined solar cells has a wide
management of the CPV systems. Distinct over height ratios of 0, 20, variety of applications in the solar power market. Accordingly, in-
40, and 60 % with inclination angles of −45°, 0° and 45° at a solar vestigation of the inclination effect on the cooling of solar cells is ne-
concentration ratios (CRs) of 5 and 10 are inspected. Based on the cessary. In this aspect and for the first time, the effects of the inclination
obtained results, the optimum inclination angle and over height ratio angle on the performance of the CPV-PCM system with zero over height
that showed the best CPV performance are identified. ratio are experimentally investigated. Fig. 8 illustrates successive pho-
tographs of the solid-liquid phase change process of RT35HC PCM at
different inclination angles of heated side 0° (vertical), 45° (heated side
3.1. Experimental results
is down), and −45° (heated side is up) when a heat flux of 3000 W/m2
is applied on the system front wall and at an ambient temperature of
The orientation of solar cells is essential in CPV systems due to the
25 ± 1 °C. Based on the figure, some important observations can be
nature of various optical light concentrators such as parabolic dish/
made. At early stages of the experiment and at all inclination angles Θ,
trough concentrators (solar irradiance is reflected by the concentrator
heat is conducted through the front aluminum plate to the adjacent
to the solar cell) and Fresnel lens concentrators (solar irradiance is
solid PCM which results in a steep rise in the temperature with time
refracted by the concentrator to the solar cell) (Browne et al., 2015).

641
R. Rabie, et al. Solar Energy 183 (2019) 632–652

Fig. 12. The liquid-solid interface evolution during melting of RT35HC at zero inclination angle (Θ = 0°) at different over height ratios of 0, 20, 40, and 60% with (I)
CR = 5 and (II) CR = 10.

until reaching the PCM melting point. At that time, a very thin layer of with continued energy input the hot buoyant liquid firstly ascends to
liquid PCM appears alongside the hot wall and the solid-liquid interface the upper part of the system, adjacent to the heated wall, then turns
is a straight vertical line as shown in Fig. 8 (t < 20 min). This behavior around and flows towards the liquid-solid interface where part of its
reveals that heat is transferred mainly through conduction during that energy is consumed for melting the solid phase. Consequently, the li-
period and in the direction normal to the hot wall. Further increase in quid becomes colder and denser then descends alongside the interface
time allows for buoyancy force to grow enough to overcome the mo- to the bottom of the container. This type of flow pattern speeds up the
lecular viscous force which allows the hot molten liquid to flow upward local heat transfer and accordingly, the rate of melting at the top por-
to the upper part of the system. At that time, buoyancy-driven con- tion of the container when compared to the lower portion, forming a
vection begins in the liquid region and the small concavity in the top manifest curvature of the melting interface (Fig. 8(b)). As melting
portion of solid-liquid interface gives a significant evidence of that continues (Fig. 8(c and d)), the solid phase, which works as a cold
(Fig. 8 (t = 20 min)). In the case of the vertical enclosure (Θ = 0°) and source driving the natural convection, begins to vanish gradually from

642
R. Rabie, et al. Solar Energy 183 (2019) 632–652

Fig. 13. The predicted vertical velocity (y-component) of the liquid PCM at 100 mm-height horizontal cross-section at two different time instants of (a) t = 35 min
and (b) t = 60 min and different over height ratios, at CR = 5.

the top part of the container allowing the molten liquid to touch the fractions rise almost linearly with time. However, as time elapses and
opposite back wall. As a result, the interface shape at the top part of the the solid PCM shrinks in the −45°-inclined system, the variation in li-
solid PCM develops a convex curvature that abates the effect of natural quid fraction deviates from a linear tendency and the melting rate re-
convection in this segment. Later, the region of weak convection cur- duces due to the poor heat transfer rate and suppression of the con-
rents extends downwards, and this trend remains until reaching com- vection current as previously mentioned. Decreasing the inclination
plete liquid phase after 150 min. angle from 0° to −45° increases the time needed to the whole melting
Increasing the inclination angle from 0° to 45° results in a higher state by 40 %. On the other hand, for the 45°-inclined system, liquid
rate of melting which is indicated by the larger melted layer thickness fraction varies almost linearly with time until the end of melting pro-
in case of the 45° inclination angle (Fig. 8(f)) when compared with that cess and achieves the highest melting rate due to the increased effects of
of the vertical orientation. The PCM liquid layer thickness at the hor- natural convection effects. The complete melting time decreases by
izontal mid plane equals 0.73 cm at the 45° inclination angle which is about 20 % when increasing the inclination angle is increased from 0°
higher than that of the vertical orientation which is 0.59 cm after to 45°. Additionally, Fig. 11 shows the unsteady variation of the
30 min from the beginning of melting process. This difference can be average hot wall temperature throughout melting of RT35HC at dif-
clarified by the fact that for the 45°-inclined system, the liquid PCM ferent inclination angles. Based on the figure, it can be deduced that the
absorbs heat from the aluminum front wall and instead of moving up- system orientation angle has significant influence on the front wall
ward parallel to the front hot wall as in the vertical orientation, it flows average temperature and consequently the CPV cell temperature. For
towards the solid-interface. This results in more heat flow from the hot instance, varying the tilting angle from 0° to 45° decreases the average
side to the solid phase attaining a higher rate of melting. Accordingly, front wall temperature by about 7 °C at a heat flux value of 3000 W/m2.
more heat transfer from the hot wall to the solid PCM and a faster On the contrary, at an angle of −45°, the average front wall tempera-
melting rate is achieved. Later, (Fig. 8(g and h)), irregular shapes are ture rises sharply with time because of the poor heat transfer rate and
detected on the solid-liquid interface in case of the 45° inclined en- the average front wall temperature is 10 °C higher than that of the
closure, which shows the existence of a 3-D flow structure and vortex vertically oriented system.
motion in the liquid phase. This structure of flow field enhances the To conclude, the system orientation angle has a remarkable effect
energy transport to the solid-liquid interface from the front hot wall, on the average temperature of the solar cell and consequently its
and the temperature uniformity of the liquid PCM. On the other hand, electrical performance. Based on the obtained results, the CPV-PCM
at Θ = −45°, the solid-liquid interface is almost linear and parallel to system with an inclination angle of 45° and 0% over height ratio ac-
the front hot wall throughout the melting process. Furthermore, the complishes the lowest average solar cell temperature due to boosting of
rate of melting is slower than that of the reference vertical angle, Θ = 0° the natural convection currents. Conversely, the highest average solar
and the small liquid layer thickness during the melting process evi- cell temperature and subsequently the lowest electrical performance
dently shows that. These observations show the presence of weak are attained at the inclination angle of −45° and 0% over height ratio.
convection in the liquid phase due to the accumulation of the hot fluid Essentially, parabolic dish/trough and Fresnel lens concentrators are
near the front hot wall which serves as an insulating layer between the usually used with CPV-PCM system with inclination angles of 45° and
liquid-solid interface and the hot wall. −45°, respectively due to the configuration of each type of con-
Accordingly, as the heat transfer rate reduces, the melting rate de- centrators. Therefore, parabolic dish/trough concentrators are more
creases with augmenting melting layer thickness. This observation preferable than lenses when using CPV-PCM system with 0 % over
agrees with (Emam et al., 2017a; Kamkari and Shokouhmand, 2014). height ratio for practical applications.
To facilitate comparisons, Fig. 9 presents the solid-liquid interface at
different time during the melting of RT35HC PCM for different tilting
3.2. Numerical results
angles at a heat flux value of 3000 W/m2 and an ambient temperature
of 25 ± 1 °C.
In this section, numerical simulations were performed to study the
For further verification, Fig. 10 shows the variation of the liquid
effects of adding a bulk of PCM over the solar cell height (hot source) on
fraction value with time during melting of RT35HC-PCM at Θ = 0°, 45°,
the cooling of the CPV systems. This is accomplished through changing
and −45°. It can be noticed that for all inclination angles, liquid
the over height ratio, Λ of the PCM cavity while keeping the PCM

643
R. Rabie, et al. Solar Energy 183 (2019) 632–652

Fig. 14. The predicted temperature fields after about 65% from the solid PCM has melted for different over height ratios of 0% to 20%, 40%, and 60% at (I) CR = 5
and (II) CR = 10.

volume and the solar cell height constant. The developed model con- system at Ʌ = 0 %, shows almost the same trends as those of the ex-
siders these modifications in three different inclination angles of −45°, perimental results at Θ = 0° and Ʌ = 0 % (Fig. 8(a–d)) and this trend is
0°, and 45°, while four different over height ratios of 0, 20, 40, and 60 % true at any of the two CRs. Secondly, increasing the over height ratio Ʌ
are investigated at each inclination angle under CR of 5 and 10. from 0 % up to 60 %, results in differences in the liquid-solid interface
shape which indicates the presence of different flow patterns in the li-
quid PCM which can be described as follows. In the case of the 0% over
3.2.1. Effects of over height ratios
height ratio, the solid PCM, which initiatives the natural convection
Fig. 12 shows the development of the liquid-solid interface during
currents, disappears progressively from the upper part of the container
melting of RT35HC at zero inclination angle (Θ = 0°) at different over
throughout the melting process as shown in Fig. 12(I) after 35 min from
height ratios of 0, 20, 40, and 60 % and CRs of 5 and 10. As seen from
the starting of melting process. Thereby, natural convection declines in
the figure, some important observations can be made. First, the esti-
this part which then extends downward with time till the complete
mated evolution of liquid–solid interface for the vertically oriented

644
R. Rabie, et al. Solar Energy 183 (2019) 632–652

Fig. 15. Ttransient average solar cell temperature at different over height ratios of 0% to 20%, 40%, and 60%, and a tilting angle of 0°, where: (a) CR = 5 and (b)
CR = 10.

Fig. 16. Local solar cell temperature after about 65% from the solid PCM has melted at different over height ratios and 0° inclination angle (a) CR = 5 and (b)
CR = 10.

melting. This situation decreases the heat transfer rate from the CPV (upward vertical component) decreases from 7 × 10−3 to 3.9 × 10−3
cell (hot source) to the solid PCM and consequently, the melting rate, m/s as time increases from 35 min to 60 min, respectively. This con-
which negatively affects the cell performance. On the other hand, firms that at 0 % over height ratio, the strength of the natural con-
varying the over height ratio, Λ from 0 % to 60 %, keeps the solid PCM vection current in this part decreases with time as the solid-phase de-
at the upper region for an extended period throughout the melting clines and its highest point shifts downwards as previously mentioned.
process. This improves natural convection currents and consequently However, at 60% over height ratio, the peak positive velocity after
the overall heat transfer process. In addition, applying different values 35 min from the start of the melting process equals about 8.5 × 10−3
of CRs on the CPV cell’s front surface results in almost identical trends m/s which is about 21% larger than that of the 0 % over height ratio at
on the solid-liquid interface evolution. However, raising the CR to 10 the same time. Moreover, this value slightly decreases from 8.5 × 10−3
leads to a noticeable intensification in the rate of melting which de- to 7.9 × 10−3 m/s as the time increases from 35 min to 60 min, re-
creases the required time to reach complete liquid phase as shown in spectively as shown in Fig. 13. These observations show that adding a
Fig. 12(II) and this trend is true at any over height ratio, Ʌ. bulk of PCM over the solar cell height via changing the over height
Further explanations are shown in Fig. 13 which compares the ratio, Λ of the PCM cavity not only increases the intensity of natural
predicted vertical velocity (y-component) of the liquid PCM at 100 mm- convection current but also maintains its strength for a prolonged
height horizontal cross-section at two-time instances of 35 min and period of time.
60 min and at different over height ratios, for CR = 5. The peak positive Fig. 14 shows the estimated temperature distribution of the PCM for
and negative velocities in the vertical component could be an indicator varied over height ratios of 0%, 20%, 40%, and 60% at CRs of 5 and 10.
of the strength of the natural convection current. As shown in the These data are presented after about 65% of the solid phase has melted
figure, it is clear that at 0 % over height ratio, the peak positive velocity for CR of 5 and 10. From Fig. 14 (I), at a CR = 5, the maximum

645
R. Rabie, et al. Solar Energy 183 (2019) 632–652

Fig. 17. The evolution of the solid-liquid interface during melting of RT35HC for different over height ratios at CR = 5 and inclination angle of 45° and −45°.

observed temperature within the PCM for the vertically oriented CPV- the molten PCM. This thermal behaviour boosts the natural convection
PCM system with a 0% over height ratio is about 71.1 °C. Furthermore, current that plays a vital role in enhancing the overall heat transfer
it can be noticed that in that case, the hot molten liquid is collected at process from the solar cell side to the solid PCM and improving tem-
the top portion of the container (adjacent to the cell) where there is no perature distribution inside the system. Increasing the CR up to 10
solid PCM and accumulates above the heavier evolving thermally- shows an identical trend to those at CR = 5, except that the rate of
stratified liquid layers. This eventually causes a non-uniform tempera- melting and the liquid PCM increases due to the increased incident solar
ture distribution along the solar cell and eventually generates hot spots. irradiance. From Fig. 14(II), the maximum observed temperature
By increasing the over height ratio Ʌ of the CPV-PCM system to 20%, within the PCM at CR = 10 is 96.6 °C, 87.7 °C, 79.8 °C, and 71.6 °C for
40%, and 60%, the maximum observed temperature within the PCM the CPV-PCM system with 0 %, 20 %, 40 %, and 60 % over height
reduced to 65 °C, 60.9 °C, and 54.5 °C, respectively. Furthermore, the ratios.
temperature of the liquid PCM is considerably reduced by increasing Fig. 15 presents the transient variation of the average cell tem-
the over height ratio and the accumulation of hot molten liquid is perature at over height ratios of 0 % to 20 %, 40 %, and 60 % at the
shifted upward and away from the solar cell. This enhancement is tilting angle of 0°. As shown in the figure, there are three different
mainly attributed to the availability of the solid PCM over the solar cell stages of temperature variation of the solar cell associated with the
height (hot source) which provides a continuous heat absorption from phase change process of the cooling material. First, a steep increase in

646
R. Rabie, et al. Solar Energy 183 (2019) 632–652

Fig. 18. The predicted temperature contours within the PCM for different over height ratios at CR = 5 inclination angle after 60 min from the starting of melting
process at inclination angle of 45° and −45°.

the average solar cell temperature followed by a gradual increase with peak at complete melting point. This variation is observed when the
time is observed. This variation is most likely due to sensible heating of solid PCM, which acts as the cold source for driving the natural con-
the PCM by conduction heat transfer through the aluminum front plate. vection currents starts to disappear from the upper part of the system.
Then, the phase transition of PCM adjacent to the aluminum front plate During this period, natural convection currents are weaker in this re-
forms a thin melting layer of the PCM while the heat transfer is still gion. Also, the temperature of the liquid PCM at the upper part of the
dominated by conduction. At this point, changing the over height ratio system increases which results in less heat transfer from the hot wall to
exhibits no difference in the average cell temperature. Secondly, further the liquid PCM. With further increase in time, the region of weak
increase in time results in a decrease in the average cell temperature convection currents extends from the top to the bottom rising the cell
and then remains almost constant for a period of time. This stage in- temperature gradually until reaching a fully liquid phase. The key idea
dicates the start of the heat transfer by convection which balances heat of adding a bulk of PCM over the solar cell height in the current re-
transfer by conduction. During that time, the natural convection current search is to delay the last stage via increasing the intensity of the nat-
is significant in the liquid PCM as the melting fraction increases. Con- ural convection current and maintaining its strength for a prolonged
sequently, the solar cell average temperature decreases. Lastly, later, period of time. At a CR = 5 (Fig. 15-a), increasing the over height ratio
the cell temperature starts to increase gradually until it reaches the of the CPV-PCM system from 0 % to 20 %, 40 %, and 60 % decreases the

647
R. Rabie, et al. Solar Energy 183 (2019) 632–652

Fig. 19. Transient variation of the average solar cell temperature for the CPV-PCM system with different over height ratios and different inclination angles (a)
Θ = 45° and (b) Θ = −45° at a CR = 5.

Fig. 20. Local solar cell temperature after about 60 min from the starting of melting process at different over height ratios and different inclination angles (a) Θ = 45°
and (b) Θ = −45° at a CR = 5.

average cell temperature from 80 °C to 77.8 °C, 76 °C, and 75 °C, re- create mechanical and thermal stresses on the cell and consequently
spectively. Increasing the CR to 10, as shown in Fig. 15b, leads to an reducing its life time. For this reason, the uniformity index of solar cell,
identical trend of the average solar cell temperature variation and this ΔT is investigated in the current study. This is indeed the difference
is true for all over height ratios. Besides, higher melting rate of PCM is between the local maximum and minimum temperature of the solar cell
shown along with raising in the average solar cell temperature. Based as shown in Eq. (15). Fig. 16 illustrates the local solar cell temperature
on the figure, significant reductions are observed in the average solar (temperature variation along the solar cell height) at variable over
cell temperatures to about 116 °C, 113 °C, 111 °C and 109 °C, by in- height ratios and CR of 5, and 10. These data are also presented after
creasing the over height ratio from 0 % to 20 %, 40 %, and 60 %, about 65 % of the solid PCM has melted for both CRs. Increasing the
respectively. over height ratio Ʌ, lowers the local temperature of the solar cell and
In addition, solar cell temperature uniformity is a key variable improves the temperature uniformity along the solar cell height and
worthy exploring to circumvent the formation of hot spots which po- this trend is true for both CRs. At CR = 5 (Fig. 16a), the uniformity
tentially causes solar cell damage. The non-uniform temperature var- index of the CPV-PCM system with 0% over height ratio is around
iation on the solar cell leads to different thermal expansion ratios which 10.2 °C. Increasing the value of the over height ratio to 20 %, 40 %, and

648
R. Rabie, et al. Solar Energy 183 (2019) 632–652

Fig. 21. Variation of the maximum temperature and temperature uniformity index versus the inclination angle of CPV-PCM system at CR = 5.

60% reduces the uniformity index to 8 °C, 7 °C and 5.9 °C, respectively. whole melting time for the −45°-inclined CPV-PCM system with 0%
An identical trend is shown in Fig. 16b at CR = 10, where the uni- and 60 % over height ratios is 129 min and 115 min, respectively.
formity index reaches 17.5 °C for the CPV-PCM system with 0% over Fig. 18 further illustrates the estimated temperature fields within
height ratio, while it is 14.2 °C, 12.7 °C and 11.3 °C for the CPV-PCM the PCM for various over height ratios at CR = 5 and inclination angles
system with 20 %, 40 %, and 60 % over height ratio, respectively. This of 45° and −45°, respectively. These data are presented after about
enhancement is mainly due to the fact that unlike in the CPV-PCM 65% of the solid PCM has melted (about 60 min from the start of
system with 0 % over height ratio, adding a bulk of PCM over the solar melting process). As expected, using an inclination angle of 45° and
cell height by increasing Ʌ reduces the cell temperature by enhancing increasing the value of the over height ratio enhances the overall heat
the heat transfer process and prevents the accumulation of hot molten transfer process in the CPV-PCM system as indicated by the reduced
liquid near to the solar cell which is shifted upwards as previously temperature of the liquid PCM. This is due to the fact that strong
shown in Figs. 14–16. This thermal behavior reduces the thermal convection currents continue till complete melting. As shown in Fig. 18,
stratification beside the cell and enhances its temperature uniformity. for the 45° oriented CPV-PCM system, the maximum observed tem-
Therefore, and based on Figs. 12–16, raising the over height ratio of perature within the PCM with 20% and 40% over height ratio is about
the vertically oriented CPV-PCM system, remarkably reduces the local 57.4 °C and 50 °C compared to 58.2 °C for the same system with 0% over
solar cell temperature and uniformity index. The CPV-PCM system with height ratio. Furthermore, increasing the over height ratio up to 60 %
a 60% over height ratio attained the best thermal performance of solar reduces the maximum temperature within the PCM to 48.1 °C. The
cell compared to the other cases. The solar cell is kept at the average figure also shows that, for the CPV-PCM system with −45° inclination
temperature of 75 °C and 109 °C while the uniformity index is 5.9 °C angle, increasing the over height ratios at the same rates decreases the
and 11.3 °C at a CR of 5 and 10, respectively. maximum observed temperature within the PCM from 83.5 °C to
72.8 °C, 63.7 °C and 57.3 °C, respectively. This reduction in the liquid
3.2.2. Effects of inclination angles PCM temperature allows more heat transfer from the cell side to the
As previously stated, the solar cells are commonly inclined in CPV PCM which reduces the cell temperature and consequently improves its
systems due to the configuration of various optical light concentrators. electrical performance. It is interesting to note that increasing the over
So, this section investigates the effects of the inclination angle on the height ratio can mitigate the negative effects of the −45° tilting angle
performance of CPV-PCM systems at different over height ratios. Fig. 17 which open doors for its use in practical applications.
presents the evolution of the solid-liquid interface during melting of Fig. 19 illustrates the variation of the transient average solar cell
RT35HC for different over height ratios at CR = 5 and inclination angle temperature for the CPV-PCM system with different over height ratios
of 45° and −45°, respectively. Based on the figure, at a 0 % over height of 0 %, 20 %, 40 %, and 60 % and various orientation angles of 45° and
ratio, the rate of melting in the 45° inclined CPV-PCM system is higher −45° at a CR = 5. For the CPV-PCM system with a 45° inclination angle
than that of the −45° -inclined CPV-PCM system. Moreover, increasing as shown in Fig. 19(a), the solar cell average temperature is kept below
the over height ratio at ether of the inclination angle raises the rate of 72.5 °C throughout the melting process and this is true for all over
melting because of the enhanced heat transfer process inside the height ratios. Moreover, increasing the over height ratio from 0% up to
system. The 45°-inclined CPV-PCM system with 60 % over height ratio 60% at the tilt angle of 45° slightly reduces the solar cell average
achieves complete liquid phase after about 95 min compared to over temperature from 72.5 °C to 70.9 °C, respectively. On the contrary, at
100 min for the same system with 0 % over height ratio. However, the the tilt angle of −45°, increasing the over height ratio remarkably

649
R. Rabie, et al. Solar Energy 183 (2019) 632–652

Fig. 22. Variation of the maximum temperature and temperature uniformity index versus the inclination angle of CPV-PCM system at CR = 10.

reduces the average temperature of the solar cell as seen in Fig. 19(b). CPV-PCM system with high values of the over height ratio should be
For instance, rising the over height ratio from 0 % up to 60 % reduces used to mitigate the negative effects of such kind of orientations.
the solar cell average temperature from 88.5 °C to 75.7 °C, respectively.
Based on these results, it is noted that the effect of increasing the over 4. Conclusion
height ratio on the solar cell average temperature is more pronounced
in case of the −45° inclined CPV-PCM system when compared to that of Experimental and numerical investigations have been carried out to
45°. Additionally, Fig. 20 shows the local solar cell temperature at investigate the influence of over height ratio of a PCM heat sink on the
different over height ratios and both tilt angles at a CR = 5. These data temperature distribution along with the solar cell. Thus, four different
are also presented after about 65% of the solid phase has melted (after over height ratios of 0 %, 20 %, 40 %, and 60 % at three inclination
about 60 min). In the case of the CPV-PCM system with 45° inclination angles of −45°, 0° and 45° with solar concentration ratios of 5 and 10
angle as shown in Fig. 20 (a), increasing the over height ratio from 0 % are investigated. In light of the current study, some important ob-
up to 60 %, decreases the uniformity index (ΔT) from 6.2 °C to 4.6 °C. servations have been made. Increasing over height ratio significantly
However, for −45° tilt angle, increasing the over height ratio from 0 % reduces the formation of stratified liquid layers, and in return reduces
to 60 %, considerably decreases the ΔT from about 14 °C to 5.3 °C, re- the potential hot spots in the upper portion of the solar cells. At a 0 %
spectively as shown in Fig. 20(b). Therefore, adding a bulk of PCM over over height ratio, employing a PCM heat sink at inclination angle of
the solar cell height of the CPV-PCM system with an orientation angle of −45° showed less cooling effect compared to inclination angles of 0°
−45°, not only reduces the average solar cell temperature but its and 45°. This behavior is mainly due to the accumulation of the liquid
temperature uniformity index as well. PCM at the top part of the PCM enclosure and near the solar cell. At the
To summarize the main findings of the current work, the effect of same time, increasing the over height ratio significantly reduces the
varying over height ratios on the temperature uniformity index at dif- solar cell maximum temperature and improves temperature uniformity
ferent CPV-PCM angles and the maximum solar cell temperatures are due to the remarkable enhancement of natural convection current in
outlined in Fig. 21(a and b), and Fig. 22 (a, and b) at CR of 5, and 10, the PCM enclosure. At an inclination angle of 45°, a slight reduction of
respectively. At CR = 5, the effect of varying over height ratio has a maximum solar cell temperature is observed with insignificant en-
favorable effect on both the maximum temperature and uniformity hancement of temperature uniformity. Accordingly, for concentrators
index. At the −45°-tilt angle, the reduction in the maximum tempera- that require a negative orientation, high values of over height ratios
ture, and uniformity index is about 17.5 °C, and 8.5 °C, respectively as should be considered.
the over height ratio varied from 0 to 60%. on the other hand, in-
creasing the inclination angle to 45 °C slightly reduces the maximum Acknowledgement
temperature and uniformity index by about 2 °C, and 1.7 °C. Similarly,
at CR = 10, at the inclination angle of −45°, the maximum cell tem- The authors would like to thank the Egyptian Ministry of Higher
perature and uniformity index are reduced by 31 °C, and 15.7 °C. Education (MoHE)-Egypt for providing a scholarship to conduct this
However, at the inclination angle of 45°, the maximum cell temperature study as well as the Egypt-Japan University of Science and Technology,
and uniformity index are deduced by 3 °C, and 1.2 °C. Thus, for con- for offering the financial support and computational tools to conduct
centrators that acquire a negative orientation such as Fresnel lenses, a this research work.

650
R. Rabie, et al. Solar Energy 183 (2019) 632–652

Appendix A

Table A-1
The specifications and accuracy of the employed instruments as reported in their datasheets.
Instrument Specifications

Silicon rubber heater Dimensions 100 mm × 100 mm


Thickness 1.5 mm
Max. temperature 250 °C
Electrical power 0.5 W/cm2
density
Heat flux sensor Dimensions 10 mm × 10 mm
Thickness 0.4 mm
Representative 0.995 μV/(W/m2)
sensitivity
Thermal resistance 0.00015 °C/(W/m2)
Accuracy ± 3%
Temperature range −200 to 200 °C
Heat flux meter Measurement range −50.00 to 50.00 mV
Accuracy ± 0.5%
Recording capacity 60,000 data
Volt slider Input voltage 100 V
Output voltage 0 to 130 V
Multichannel temperature Number of channels 4
recorder Measurement range −270 to 1370 °C
Accuracy ± 0.5 °C
Resolution 0.1 °C
Recording capacity 240,000 data/ch.
Recording interval 0.1 s to 30 min
Micro-thermocouples Type K-type
Covering material Teflon
Wire diameter 0.127 mm (Temp. sensor) and
0.32 mm (connector)
Temperature range 0 to 200 °C
Insulation Thickness 30 mm
Thermal k = 0.026 W/m K
conductivity

References material heat sink for low concentrated photovoltaic systems. In: ASME International
Mechanical Engineering Congress and Exposition, Proceedings (IMECE). https://doi.
org/10.1115/IMECE2017-70937.
Agrawal, S., Tiwari, G.N., 2011. Energy and exergy analysis of hybrid micro-channel Green, M.A., 2013. Silicon solar cells: state of the art. Philos. Trans. A. Math. Phys. Eng.
photovoltaic thermal module. Sol. Energy 85, 356–370. https://doi.org/10.1016/j. Sci. 371, 20110413. https://doi.org/10.1098/rsta.2011.0413.
solener.2010.11.013. Hasan, A., McCormack, S.J., Huang, M.J., Norton, B., 2010. Evaluation of phase change
Ahmed, M., Radwan, A., 2017. Performance evaluation of new modified low-concentrator materials for thermal regulation enhancement of building integrated photovoltaics.
polycrystalline silicon photovoltaic/thermal systems. Energy Convers. Manage. 149, Sol. Energy 84, 1601–1612. https://doi.org/10.1016/j.solener.2010.06.010.
593–607. https://doi.org/10.1016/j.enconman.2017.07.057. Hasan, A., Sarwar, J., Alnoman, H., Abdelbaqi, S., 2017. Yearly energy performance of a
AS ONE [WWW Document], 2018. Tech. data sheet. URL < https://www.as-1.co.jp/en/ photovoltaic-phase change material (PV-PCM) system in hot climate. Sol. Energy
merchandise/ > . 146, 417–429. https://doi.org/10.1016/j.solener.2017.01.070.
Atkin, P., Farid, M.M., 2015. Improving the efficiency of photovoltaic cells using PCM Hedayatizadeh, M., Ajabshirchi, Y., Sarhaddi, F., Safavinejad, A., Farahat, S., Chaji, H.,
infused graphite and aluminium fins. Sol. Energy 114, 217–228. https://doi.org/10. 2013. Thermal and electrical assessment of an integrated solar photovoltaic thermal
1016/j.solener.2015.01.037. (PV/T) water collector equipped with a compound parabolic concentrator (CPC). Int.
Brent, A.D., Voller, V.R., Reid, K.J., 1988. Enthalpy-porosity technique for modeling J. Green Energy 10, 494–522. https://doi.org/10.1080/15435075.2012.678524.
convection-diffusion phase change: application to the melting of a pure metal. Horowitz, K.A.W., Fu, R., Sun, X., Silverman, T., Woodhouse, M., Alam, M.A., Horowitz,
Numer. Heat Transf. 13, 297–318. https://doi.org/10.1080/10407788808913615. K.A.W., Fu, R., Sun, X., Silverman, T., Woodhouse, M., Alam, A., 2017. An analysis of
Browne, M.C., Norton, B., McCormack, S.J., 2015. Phase change materials for photo- the cost and performance of photovoltaic systems as a function of module area,
voltaic thermal management. Renew. Sustain. Energy Rev. 47, 762–782. https://doi. National Renewable Energy Laboratory (NREL).
org/10.1016/j.rser.2015.03.050. Huang, M.J., Eames, P.C., Norton, B., Hewitt, N.J., 2011. Natural convection in an in-
Captec – heat flux sensor [WWW Document], 2018. URL < https://www.captec.fr/copie- ternally finned phase change material heat sink for the thermal management of
de-fluxmetre-thermique > . photovoltaics. Sol. Energy Mater. Sol. Cells 95, 1598–1603. https://doi.org/10.1016/
Dhaidan, N.S., Khodadadi, J.M., Al-Hattab, T.a., Al-Mashat, S.M., 2013. Experimental and j.solmat.2011.01.008.
numerical investigation of melting of phase change material/nanoparticle suspen- Kamkari, B., Amlashi, H.J., 2017. Numerical simulation and experimental verification of
sions in a square container subjected to a constant heat flux. Int. J. Heat Mass Transf. constrained melting of phase change material in inclined rectangular enclosures. Int.
66, 672–683. https://doi.org/10.1016/j.ijheatmasstransfer.2013.06.057. Commun. Heat Mass Transf. 88, 211–219. https://doi.org/10.1016/j.
Dong, J., Zhuang, X., Xu, X., Miao, Z., Xu, B., 2018. Numerical analysis of a multi-channel icheatmasstransfer.2017.07.023.
active cooling system for densely packed concentrating photovoltaic cells. Energy Kamkari, B., Groulx, D., 2018. Experimental investigation of melting behaviour of phase
Convers. Manage. 161, 172–181. https://doi.org/10.1016/j.enconman.2018.01.081. change material in finned rectangular enclosures under different inclination angles.
Emam, M., Ahmed, M., 2018a. Cooling concentrator photovoltaic systems using various Exp. Therm. Fluid Sci. 97, 94–108. https://doi.org/10.1016/j.expthermflusci.2018.
configurations of phase-change material heat sinks. Energy Convers. Manage. 158, 04.007.
298–314. https://doi.org/10.1016/j.enconman.2017.12.077. Kamkari, B., Shokouhmand, H., 2014. Experimental investigation of phase change ma-
Emam, M., Ahmed, M., 2018b. Performance analysis of a new concentrator photovoltaic terial melting in rectangular enclosures with horizontal partial fins. Int. J. Heat Mass
system integrated with phase change material and water jacket. Sol. Energy 173, Transf. 78, 839–851. https://doi.org/10.1016/j.ijheatmasstransfer.2014.07.056.
1158–1172. https://doi.org/10.1016/j.solener.2018.08.069. Kamkari, B., Shokouhmand, H., Bruno, F., 2014. Experimental investigation of the effect
Emam, M., Ookawara, S., Ahmed, M., 2017a. Performance study and analysis of an in- of inclination angle on convection-driven melting of phase change material in a
clined concentrated photovoltaic-phase change material system. Sol. Energy 150, rectangular enclosure. Int. J. Heat Mass Transf. 72, 186–200. https://doi.org/10.
229–245. https://doi.org/10.1016/j.solener.2017.04.050. 1016/j.ijheatmasstransfer.2014.01.014.
Emam, M., Radwan, A., Ahmed, M., 2017b. Analysis of a new hybrid water-phase change Kant, K., Shukla, A., Sharma, A., Biwole, P.H., 2016a. Heat transfer studies of

651
R. Rabie, et al. Solar Energy 183 (2019) 632–652

photovoltaic panel coupled with phase change material. Sol. Energy 140, 151–161. thermal (PV/T) collector. Renew. Energy 77, 43–50. https://doi.org/10.1016/j.
https://doi.org/10.1016/j.solener.2016.11.006. renene.2014.12.012.
Kant, K., Shukla, A., Sharma, A., Biwole, P.H., 2016b. Thermal response of poly-crys- Rejeb, O., Sardarabadi, M., Menezo, C., Passandideh-Fard, M., Dhaou, M.H., Jemni, A.,
talline silicon photovoltaic panels: numerical simulation and experimental study. Sol. 2016. Numerical and model validation of uncovered nanofluid sheet and tube type
Energy 134, 147–155. https://doi.org/10.1016/j.solener.2016.05.002. photovoltaic thermal solar system. Energy Convers. Manage. 110, 367–377. https://
Khanna, S., Reddy, K.S., Mallick, T.K., 2018. Climatic behaviour of solar photovoltaic doi.org/10.1016/j.enconman.2015.11.063.
integrated with phase change material. Energy Convers. Manage. 166, 590–601. Royne, A., Dey, C.J., Mills, D.R., Royane, A., Dey, C.J., Milis, D.R., Royne, A., Dey, C.J.,
https://doi.org/10.1016/j.enconman.2018.04.056. Mills, D.R., 2005. Cooling of photovoltaic cells under concentrated illumination: a
Nizetic, S., Coko, D., Yadav, A., Grubisic-Cabo, F., 2016. Water spray cooling technique critical review. Sol. Energy Mater. Sol. Cells 86, 451–483. https://doi.org/10.1016/j.
applied on a photovoltaic panel: the performance response. Energy Convers. Manage. solmat.2004.09.003.
108, 287–296. https://doi.org/10.1016/j.enconman.2015.10.079. Rubitherm GmbH [WWW Document], 2014. Tech. data sheet. URL Rubitherm
Pal, D., Joshi, Y.K., 2001. Melting in a side heated tall enclosure by a uniformly dis- Technologies GmbH Berlin, Germany.
sipating heat source. Int. J. Heat Mass Transf. 44, 375–387. https://doi.org/10.1016/ Sarwar, J., Georgakis, G., LaChance, R., Ozalp, N., 2014. Description and characterization
s0017-9310(00)00116-2. of an adjustable flux solar simulator for solar thermal, thermochemical and photo-
Rabie, R., Hassan, H., Ookawara, S., Ahmed, M., 2017. Performance enhancement of the voltaic applications. Sol. Energy 100, 179–194. https://doi.org/10.1016/j.solener.
concentrated photovoltaic using different phase change material configurations. 2013.12.008.
Energy Procedia 141, 61–65. https://doi.org/10.1016/j.egypro.2017.11.012. Sharma, S., Micheli, L., Chang, W., Tahir, A.A., Reddy, K.S., Mallick, T.K., 2017. Nano-
Radwan, A., Ahmed, M., 2018. Thermal management of concentrator photovoltaic sys- enhanced Phase Change Material for thermal management of BICPV. Appl. Energy
tems using microchannel heat sink with nanofluids. Sol. Energy 171, 229–246. 208, 719–733. https://doi.org/10.1016/j.apenergy.2017.09.076.
https://doi.org/10.1016/j.solener.2018.06.083. Soliman, A.M.A., Hassan, H., 2019. Effect of heat spreader size, microchannel config-
Radwan, A., Ookawara, S., Ahmed, M., 2019. Thermal management of concentrator uration and nanoparticles on the performance of PV-heat spreader-microchannels
photovoltaic systems using two-phase flow boiling in double-layer microchannel heat system. Sol. Energy 182, 286–297. https://doi.org/10.1016/j.solener.2019.02.059.
sinks. Appl. Energy, accepted. https://doi.org/10.1016/j.apenergy.2019.03.017. Stropnik, R., Stritih, U., 2016. Increasing the efficiency of PV panel with the use of PCM.
Radwan, A., Ahmed, M., 2017. The influence of microchannel heat sink configurations on Renew. Energy 97, 671–679. https://doi.org/10.1016/j.renene.2016.06.011.
the performance of low concentrator photovoltaic systems. Appl. Energy 206, Xu, Z., Kleinstreuer, C., 2014. Concentration photovoltaic-thermal energy co-generation
594–611. https://doi.org/10.1016/j.apenergy.2017.08.202. system using nanofluids for cooling and heating. Energy Convers. Manage. 87,
Radwan, A., Ahmed, M., Ookawara, S., 2016a. Performance enhancement of concentrated 504–512. https://doi.org/10.1016/j.enconman.2014.07.047.
photovoltaic systems using a microchannel heat sink with nanofluids. Energy Zarma, I., Ahmed, M., Ookawara, S., 2019. Enhancing the performance of concentrator
Convers. Manage. 119, 289–303. https://doi.org/10.1016/j.enconman.2016.04.045. photovoltaic systems using nanoparticle-phase change material heat sinks. Energy
Radwan, A., Emam, M., Ahmed, M., 2017. Comparative study of active and passive Convers. Manage. 179, 229–242. https://doi.org/10.1016/j.enconman.2018.10.055.
cooling techniques for concentrated photovoltaic systems. In: Exergetic, Energetic Zeng, L., Lu, J., Li, Y., Li, W., Liu, S., Zhu, J., 2017. Numerical study of the influences of
and Environmental Dimensions. Elsevier, pp. 475–505. https://doi.org/10.1016/ geometry orientation on phase change material’s melting process. Adv. Mech. Eng. 9,
B978-0-12-813734-5.00027-5. 1–11. https://doi.org/10.1177/1687814017720084.
Radwan, A., Ookawara, S., Ahmed, M., 2016b. Analysis and simulation of concentrating Zhou, J., Yi, Q., Wang, Y., Ye, Z., 2015. Temperature distribution of photovoltaic module
photovoltaic systems with a microchannel heat sink. Sol. Energy 136, 35–48. https:// based on finite element simulation. Sol. Energy 111, 97–103. https://doi.org/10.
doi.org/10.1016/j.solener.2016.06.070. 1016/j.solener.2014.10.040.
Radwan, A., Ookawara, S., Mori, S., Ahmed, M., 2018. Uniform cooling for concentrator Zhu, Z.Q., Huang, Y.K., Hu, N., Zeng, Y., Fan, L.W., 2018. Transient performance of a
photovoltaic cells and electronic chips by forced convective boiling in 3D-printed PCM-based heat sink with a partially filled metal foam: effects of the filling height
monolithic double-layer microchannel heat sink. Energy Convers. Manage. 166, ratio. Appl. Therm. Eng. 128, 966–972. https://doi.org/10.1016/j.applthermaleng.
356–371. https://doi.org/10.1016/j.enconman.2018.04.037. 2017.09.047.
Rejeb, O., Dhaou, H., Jemni, A., 2015. A numerical investigation of a photovoltaic

652

You might also like