You are on page 1of 16

Solar Energy 179 (2019) 135–150

Contents lists available at ScienceDirect

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

Numerical and outdoor real time experimental investigation of performance T


of PCM based PVT system
H. Fayaza,b, N.A. Rahima,c, M. Hasanuzzamana,b, , A. Rivaia, R. Nasrind

a
UM Power Energy Dedicated Advanced Centre (UMPEDAC), Level 4, Wisma R&D, University of Malaya, 59990 Kuala Lumpur, Malaysia
b
Institute for Advanced Studies, University of Malaya, 50603 Kuala Lumpur, Malaysia
c
Renewable Energy Research Group, King Abdulaziz University, Jeddah 21589, Saudi Arabia
d
Department of Mathematics, Bangladesh University of Engineering and Technology, Dhaka 1000, Bangladesh

ARTICLE INFO ABSTRACT

Keywords: Photovoltaic power generation is a suitable option to counter depleting and environmentally hazardous fossil
Solar irradiation fuels. However, increased cell temperature of the photovoltaic module reduces the electrical performance.
PV/T Therefore, for enhancing the electrical performance as well as to obtain the useful thermal, a combined pho-
PCM tovoltaic thermal system is suitable technology. Furthermore, the addition of phase change materials into
Performance
photovoltaic thermal systems adds more dual benefits in terms of cooling of PV cell as well as heat storage.
Hence, there are still issues to transfer heat from the system efficiently, which cause lower performance of PVT
and PVT-PCM systems. In this paper, the aluminium material of thermal collector is used by introducing a novel
design to enhance heat transfer performance, which is assembled in PVT and PVT-PCM systems. Experimental
validation is carried out for the 3D FEM-based numerical analysis with COMSOL Multiphysics® at 200 W/m2 to
1000 W/m2 varying irradiation levels while keeping mass flow rate fixed at 0.5LPM and inlet water temperature
at 32 °C. The experiment is carried out at outdoor free weather conditions with passive cooling of the module by
an overhead water tank scheme. A good agreement in numerical and experimental results is achieved through
experimental validation. Cell temperature reduction of 12.6 °C and 10.3 °C is achieved from the PV module in
case of the PVT-PCM system. The highest value of the electrical efficiency achieved is 13.72 13.56% for PV and
13.85 and 13.74% for PVT numerically and experimentally respectively. Similarly, for PVT-PCM, electrical ef-
ficiency is achieved as 13.98 and 13.87% numerically and experimentally respectively. In the case of the PVT
system, electrical performance is improved as 6.2 and 4.8% and for PVT-PCM, it is improved as 7.2 and 7.6% for
numerically and experimentally respectively.

1. Introduction economically feasible as compared to separate systems (Al-Waeli et al.,


2018c). These both systems are often used separately. The PV cells
Climate change and depleting conventional energy sources with convert solar energy with considerable less percentage of total incident
increasing energy demand has created concerns amongst the govern- irradiations (normally between 10 and 20%). The cells are very ab-
ments and research all over the world. Therefore, renewable energy sorbent towards all solar radiation, which causes the major part of solar
sources are being studied harnessed to combat the issues stated (Ahmed radiations to transform into heat, making the cells hotter and as a re-
et al., 2013; Fayaz et al., 2011). Renewable energy sources provide a sult, which reduces its electrical efficiency (Chauhan et al., 2018). Cell
sustainable alternative to our dependence on fossil fuels and reduce temperature is very important regarding cell efficiency even though the
carbon emissions. Hence, it is necessary to develop an innovative other operating conditions be set at optimum for improved performance
technique of solar energy conversion as a better alternative to where because the electrical performance of PV module declines due to en-
the use of fossil fuels is at the highest risk. Solar energy is one of the ormous cell temperature gain. Observation of electrical efficiency of
abundant and promising available renewable sources. There are two 0.469% is done with each degree Celsius drop in PV cells (Lim et al.,
uses of solar energy such as thermal and electrical conversions. Thermal 2013; Skoplaki and Palyvos, 2009). Nishioka et al. (2003) installed a PV
collectors produce thermal energy and photovoltaic produce electrical system with a capacity of 50 kW in Japan for its field testing. It was
energy, which allows both advantages in one system making it observed that the PV system is affected by high temperature when being


Corresponding author at: UM Power Energy Dedicated Advanced Centre (UMPEDAC), Level 4, Wisma R&D, University of Malaya, 59990 Kuala Lumpur, Malaysia.
E-mail address: hasan@um.edu.my (M. Hasanuzzaman).

https://doi.org/10.1016/j.solener.2018.12.057
Received 24 July 2018; Received in revised form 19 November 2018; Accepted 23 December 2018
0038-092X/ © 2018 Elsevier Ltd. All rights reserved.
H. Fayaz et al. Solar Energy 179 (2019) 135–150

Nomenclature Tb Tedlar back surface temperature (°C)


Tg Glass surface temperature (°C)
A Area of PV surface (m2) Tr Reference temperature (°C)
B Bias error Ts Sky temperature (°C)
Cp Specific heat at constant pressure (J kg−1 K−1) Tsc Temperature of solar cell (°C)
ch Channel Uc Measured uncertainty
Ee Module’s electrical energy (W) Usca Coefficient of heat transfer through glass from top layer to
Ein Incident energy at module top surface (W) ambient (Wm−2 K−1)
Et Useful thermal energy in the system (W) Ut Coefficient of heat transfer from top layer to bottom tedlar
El Lost energy from glass layer to ambient (W) layer (Wm−2 K−1)
f Fluid Ub Coefficient of heat transfer from tedlar layer to ambient
fu Fusion (Wm−2 K−1)
G Solar irradiance (W m−2) W Width of bottom tedlar surface (m)
he Heat exchanger
K Thermal conductivity (Wm−1 K−1) Greek symbols
Nu Rate of heat transfer at bottom surface
l Liquid αb Absorptivity of tedlar back sheet
L Latent heat αg Absorptivity of glass
N Number of data αsc Absorptivity of PV module
PV Photovoltaic σ (=5.670367 × 10−8 W m−2 K−4) Stefan- Boltzmann
Psc Packing factor of module constant
pcm Phase change material τg Transmitivity of glass
Q Thermal energy (W) εg Emissivity of glass
tλ Estimate of precision error ηsc Reference electrical efficiency of PV module
s Solid ρ Density (kg m−3)
SF Standard deviation μsc PV temperature coefficient (%/°C)
T Fluid temperature (°C) θ liquid phase fraction
Ta Ambient temperature (°C) λ Degree of freedom

used under a vast temperature range. Every time the temperature performance of PVT systems in terms of electrical and thermal effi-
coefficient rises by 0.1%/°C, about a 1% increase of annual output ciencies (Hasanuzzaman et al., 2015; Riffat and Cuce, 2011). Its per-
energy happens because of the temperature coefficient on conversion formance also relies on climate conditions. Standard test condition
efficiency. (STC) differ according to the type of PV cell technology and the climate
Combined systems comprising PV and thermal collector together are conditions they are under (Daghigh et al., 2011; Vokas et al., 2006).
on top of the priority for the reason that thermal energy is obtained Amorphous silicon PV is found to perform well under Malaysia’s cli-
while reducing cell temperature. Such combination of the collectors mate condition in comparison to other types of PV cells (Adiyabat et al.,
causes the increase in the cell electrical efficiency along with reduced 2006). Browne et al. (2015) conducted the performance investigation of
sized of solar collectors (Al-Waeli et al., 2016; Bertram et al., 2012; the PVT-PCM system at controlled working conditions at the indoor
Dupeyrat et al., 2014). There are many applications of phase change environment. Under simulated conditions, PVT-PCM systems stored
materials such as battery thermal management, thermal energy storage, more heat as compared to the PVT system with 6 °C more temperature
smart thermal fibres and photovoltaic thermal management etc. due to was gained by water in PVT-PCM. However, the PVT-PCM system took
their thermal characteristics. Recently, PCMs are used to regulate the double time to store heat energy as compared to the PVT system. Heat
thermal energy in photovoltaic thermal systems to get the better elec- storage potential increased by 100% in the higher temperature climate.
trical performance and thermal energy (Song et al., 2018). Phase The PVT-PCM system’s amount of heat is shown to be increased as
change materials (PCMs) incorporated into PVT systems provide more compared to Irish maritime climate. Furthermore, the PVT-PCM system
advantage in storing heat as well as reducing more temperatures of PV stored the heat for double-timing as compared to the PVT system for an
cell as compared to PVT systems without the use of PCMs. The latent extended time of use. Simulation of one dimensional energy balance
heat of phase change materials is quite higher which let the material model is carried out for PVT-PCM by Malvi et al. (2011). It is found that
store heat when the transition temperature of materials reaches from with the use of PCM, the performance of PV can be increased 9% along
solid to liquid. Whereas, it releases latent heat in the cooler environ- with average water temperature rise of 20 °C.
ment to surroundings while turning from liquid to solid form. However, In this paper, the effect of solar irradiations of different intensities
PCM stores sensible heat as well as a constant phase of solid or liquid on PV, PVT and PVT-PCM systems is investigated at outdoor weather
(Browne et al., 2015; Nouira and Sammouda, 2018). Phase change conditions. Cell temperature, temperature distribution over the whole
materials have different values of thermal conductivity which is an system, thermal energy collection, output electrical power and overall
important factor for heat transfer form PV cell to PCM. However, it can performance of the PV, PVT and PVT-PCM systems have been in-
be enhanced by introducing nano particles into PCMs making it as vestigated and compared.
nano-PCM (Aberoumand et al., 2018; Al-Waeli et al., 2018a; 2018b; The novelty of this work is 3D numerical analysis is carried out in
Chauhan et al., 2018; Ebadi et al., 2018; Mousavi et al., 2018). In ad- place of 1D analysis for detail investigation of heat transfer from each
dition, thermal performance can further be enhanced by using nano- layer of PV module and to the thermal collector used with water as
fluid as working fluid (Al-Waeli et al., 2017; Rostami et al., 2018). Wolf working fluid. 3D numerical analysis is rarely done for such huge sys-
(1976) discovered in his investigation on PVT collectors that they can tems where it gives critical analysis and data to further understand the
produce hot water by using Hottel-Whillier model (Hottel and Whillier, heat transfer phenomenon in better way. The new thermal collector
1955), which is an economically viable method. The design of the design using aluminum as its material with maximum thermal collector
system and operating conditions are primary aspects that affect the length to target maximum output temperature from the PV module is

136
H. Fayaz et al. Solar Energy 179 (2019) 135–150

used. A parallel photovoltaic thermal system using PCM (paraffin k 2T 2T 2T

group) is used for the temperature transition of 44 °C) for the weather of + + =0
Cp x2 y2 z2 (11)
Malaysian conditions for maximum electrical efficiency gain.
Furthermore, the experimental validation of the model is carried out For the fluid domain
extensively with set working conditions which have proved the math- u v w
ematical model with approximate allowance of 5% error percentage in + + =0
x y z (12)
the experimental results.
ui u u p 2u 2u 2u

2. Research methods u +v i +w i = +µ i
+ i
+ i
x y z xj x2 y2 z2 (13)

2.1. Numerical investigation T T T 2T 2T 2T


( Cp ) u +v +w =k + +
x y z x2 y2 z2 (14)
The solar energy total amount received by PV, PVT and PVT-PCM is
calculated as (Nishioka et al., 2003) where i = j = 1, 2, 3 represents u, v and w component of the velocity
vector of fluid and x, y, z-direction respectively.
Er = g sc psc GA (1) µf Cpf Vin Dh f
Pr = kf
is the Prandtl number and Re = µf
is the Reynolds
Some of the energy is dissipated back into the environment, which is number where Dh is the hydraulic diameter, represented by the equa-
given as; 4A
tion Dh = p in . Here Ain, pin, ρf, Vi, µ f , Cp f and kf, are an area of the
in
El = Usc (Tsc Tamb ) A (2) section of inlet boundary, inlet pipe perimeter, density, inlet fluid ve-
locity, dynamic viscosity, specific heat at a consistent pressure and
Electrical power is produced from the absorbed energy on the solar
thermal conductivity of fluid respectively.
cell as the following equation (Nasrin et al., 2018a);
Ep = sc psc g sc GA [1 µsc (Tsc Tr )] (3) 2.1.2. Numerical procedure for the PVT-PCM system
When PCMs change their phase, the heat is transferred during this
Further, the rest energy is transferred through conduction process
process. This model describes the heat transformer phenomenon in
form cell to Tedlar of PV module, which is given in Eq. (4).
PCMs assembled in the geometry as shown in Figs. 3a and 3b. For the
Et = Ut (Tsc Ttd ) A (4) numerical simulation of the PVT-PCM collector, enthalpy-based method
Following the energy balance equation shows the energy from top formulation is used. A phase change zone is attached to the flow
glass to Tedlar and environment. channel of the PVT system with thermal paste by using this method.
Convection in the liquid phase is not considered in this model because it
Er = El + Et + Ee (5) not significant during liquid forms of PCM in this case.
Solar cell temperature can be calculated from the Eq. (6) (Rahman Assumptions made for this numerical investigation is given below;
et al., 2015).
a) Phase change based volume difference is ignored.
psc G ( sc ) + (Usca Ta + Ut Ttd)
Tsc =
g sc
b) In both cases of solid and liquid, the phase is isotropic and homo-
(Usca + Ut ) (6) geneous as well as the segmental melting process is symmetric.
The extracted thermal energy from the PV module by the fluid can c) The solidification temperature is constant, and subcooling effects
be obtained by Eq. (7). are ignored due to the assumption of ideal solidification behaviour
of PCM.
Et = mCpf (Tout Tin) (7) d) Homogeneous distribution of solid in the mushy region is considered
PV module electrical efficiency can be calculated as; e) It is assumed that contact between the flow channel and PCM bags if
perfect.
produced electrical power Ep
e = = 2T 2T 2T
Total received energy Ein (8) kpcm, s
pcm, s
+
pcm, s
+
pcm, s
= Uch (Tch Tpcm, s )
x2 y2 z2 (15)
Eq. (9) is for thermal efficiency of said systems.
2T 2T 2T
Extracted thermal energy E pcm, l pcm, l pcm, l
t = = t kpcm, l + + = Upcm (Tpcm, l Tch )
Total received energy Ein (9) x2 y2 z2 (16)
Eq. (10) is for overall efficiency Initially, the PCM is in the solid phase. For the presence of solid
Ep + Et irradiation, PCM tries to collect heat from the flow channel and melting
= procedure starts as followed by Eq. (13). Following equation is defined
o
Ein (10)
considering liquid fraction values, θ, for melting PCM for modelling
properties changes between the two phases of PCM, (Eq. (15)).
2.1.1. Numerical procedure for PVT system 1
The 3D numerical simulation has experimented under regular con- pcm, s = l + (1 ) s ; Cp, pcm, s = [ l Cp, l + (1 ) s Cp,s] ; k pcm, s
pcm, s
ditions. Heat flow is investigated from each layer of the PVT system.
= kl + (1 ) ks (17)
There are some test conditions set as it is the incompressible laminar
flow and temperature inconsistencies through the thickness are ig- The thermal conductivity, density and specific heat capacity of PCM
nored, surfaces are free of any dust that will disrupt solar energy ab- in solid and liquid forms are, ks, kl, ρs, ρl, Cp,s, Cp,l respectively, in the
sorptivity, and transmission quality is 100% approximately for ethyl above equations. The mass of liquid and solid ratio to the total mass of
vinyl acetate (EVA). There is no role of wind velocity in this study. PCM called as a liquid, and solid fractions are between 0 and 1 value,
Below are the ruling partial differential equations for the thermal en- which are determined at every iteration of the total solution. When
ergy of different stratum along with equations of continuity and mo- PCM is in total liquid from the liquid fraction is set to 1, and when the
mentum of the system (Fayaz et al., 2018; Nasrin et al., 2018a, 2018b): PCM is in total solid form the liquid the fraction is set to 0, similar
For the solid layers procedure happens for the solid procedure.

137
H. Fayaz et al. Solar Energy 179 (2019) 135–150

In this model, the definition of solid fraction ϕ and liquid fraction θ


in COMSOL is done by inserting a step function which changes between
0 and 1. This step function is dependent on the changes in temperatures
around PCM phase change temperature. Furthermore, a variation of 1 K
up and down to phase change temperature of the material is set a limit
for the phase change zone. (a) PV
At the same time, by appending the latent heat of fusion impact as
well as the latent heat of solidification, values should be modified due
to phase change phenomenon. Therefore, for adding a normalised pulse
parameter, D (K−1), an analytic function is defined in COMSOL as a
coefficient multiplied by latent heat of fusion for both solid and liquid
phases; latent heat for solidification for both phases of solid and liquid.
Notably, over the temperature range between the liquid and solid
phases, the integral of D(T) must be equal to unity.
Cp, l = Cp, l, phase + DLfu ; Cp, s = Cp, s, phase + DLfu (b) PVT

Cp, s = Cp, s, phase + DLso ; Cp, l = Cp, l, phase + DLso

where Lfu, Lso, Cp,l,phase and Cp,s,phase the latent heat of fusion, latent heat
of solidification, specific heat capacities of the liquid and solid phases
respectively.

2.1.2.1. Boundary conditions.

Tg
1. At the top surface of the PV: inward heat flux k g z = G
2. At the top surface of the PV: diffuse surface condition (c) PVT-PCM
n. q = (Tamb 4
Tg4) , where σ (=5.670367 × 10−8 W m−2 K−4)
is Stefan-Boltzmann constant.
3. At the top surface of the PV: convective heat loss Fig. 1. Generation of finite element mesh for PV, PVT and PVT-PCM module.
n. q = Usca (Tamb Tg )
4. At the solid-fluid interface: kf ( )T
n f
= khe ( )
T
n he
considered is the temperature of the solar cell. No significant change in
the mean value of cell temperature is observed at the expense of un-
5. At all solid boundaries of the fluid passing pipe: no-slip condition
bearable time length as depicted in a third and fourth column of
u=v=w=0
Table 1. Therefore, for numerical simulations, values in the third
6. At the inlet boundary of the aluminum pipe: T = Tin , u = Uin, v = 0,
column for PV, PVT and PVT-PCM are taken into consideration.
w=0
7. At the outlet boundary: convective boundary condition p = 0
T
8. At all other boundaries of PVT module: insulation ns = 0 2.3. Experimental setup

where n is the distance running along x or y or z directions re- An experimental setup is prepared to acquire data at Level-3 Wisma
sponding normally to the surface. R&D outdoor solar testing facility, University of Malaya. The experi-
ment is carried out under free weather conditions of the average 32 °C
2.2. Numerical technique environmental and water inlet temperature. The experimental setup
mainly comprises the solar simulator, PV, PVT and PVT-PCM systems,
To solve the numerical model, COMSOL Multiphysics is used to which are assembled at the UMPEDAC mechanical facility. Indoor
apply the finite element method (FEM) in this study. The method schematic experimental setup is shown in Fig. 2.
(Chandrasekar et al., 2013) is utilized to solve the system of Equations The onsite instruments have taken necessary data; I-V tracer is used
used for the current experiment. To understand the heat distribution of for measuring all the essential electrical parameters of the systems.
solar cells and output fluid (Tsc, Tout) and velocity distribution of the Standard instrumentation takes meteorological data, solar irradiation
fluids (u, v, w) of the PVT module, the conservation of mass and energy and other electrical and thermal parameters. Data Taker DT80 model
and momentum equations are imperative factors in this study. Data data logger is used for data logging in 60 s interval. Instruments used
from COMSOL Multiphysics is recorded in the form of graphs in Mi- are; P1024 model silicon pyranometer for irradiations, LZM-15 flow
crosoft Excel based on the data. meter for mass flow rate and K-type thermocouples for temperature
data. Tedlar temperature is important to obtain for final cell tempera-
2.2.1. Mesh generation and grid check ture assessment. Therefore, one thermocouple is attached with Tedlar.
Fig. 1 shows the computational domain meshing of three systems However, another thermocouple is connected to the PV module surface
PV, PVT, and PVT-PCM. Finer type and coarse type meshing’s chosen to understand the surface temperature accumulation. Similarly, surface
for these numerical simulations of PV, PVT and PVT-PCM modules re- and Tedlar thermocouples are attached to PVT and PVT-PCM systems
spectively from the built-in-physics-controlled meshing. Three-dimen- along with fluid inlet and outlet of the systems. All the relevant data
sional (3D) model is chosen for numerical analysis. For the subdomains such as temperatures, photovoltaic current, voltage, output power,
and boundaries of the numerical model, free tetrahedral and triangular ambient temperatures, humidity, wind speed and solar irradiation were
elements are used respectively. At conditions of 1000 W/m2 solar ir- taken. Schematic cross-sectional view of PVT system and a plan view of
radiations and 0.5 LPM of mass flow rates, two tests independent of the the thermal collector are shown in Figs. 3a and 3b. Fig. 4 shows the
grid are performed for PV, PVT and PVT-PCM systems. experimental fabrication process of PVT and PVT-PCM system along
For PV, PVT and PVT-PCM systems, different non-uniform grid with phase change material used in the system. Two separate PVT and
systems are checked as shown in Table 1. The supervising parameter PVT-PCM systems are prepared.

138
H. Fayaz et al. Solar Energy 179 (2019) 135–150

Table 1
Grid sensitivity check at 1000 W/m2 irradiation.
PV Type of meshing Normal Fine Finer Extra fine
Elements 59,636 1,87,601 11,66,206 79,50,780
Cell temperature (°C) 69.52532 75.12827 75.12881 75.12890
Time of solution (s) 26 69 789 5028

PVT Type of meshing Extra coarse Coarser Coarse Normal


Elements 14,99,353 20,33,676 36,64,990 74,00,382
Cell temperature (°C) 67.00529 67.00559 67.00582 67.00593
Time of solution (s) 6650 9102 15,021 31,541

PVT-PCM Type of meshing Extra coarse Coarser Coarse Normal


Elements 15,51,566 20,98,558 37,09,748 78,54,851
Cell temperature (°C) 63.01413 63.01458 63.01479 63.01491
Time of solution (s) 6918 9546 16,379 32,932

Fig. 2. Schematic experimental setup of the PVT system.

Fig. 3a. Cross-sectional view diagram of the photovoltaic thermal system.

The data taken has been analysed to identify the effect of different taken is for a long period to get consistent irradiation levels and other
solar irradiation effects on cell temperature, temperature distribution weather conditions. The outdoor experimental setup is shown in Fig. 5.
into the whole system, thermal energy collection, output electrical Manufactured in Malaysia, PV module model is E310P(S)-011 with
energy and overall performance of the PV, PVT and PVT. The mass flow a solar cell size of 156 × 156 mm comprising total 72 cells making the
rate is fixed at 0.5 LPM which is laminar with Reynolds number less overall PV module size of 2 m2. Table 2 shows the further technical
than 2000. The data is collected during the whole day from about 8:30 details of the PV module used to experiment.
am to 7:00 pm from May to August 2017. The solar irradiations There are five layers of the PV module used, assembled in the me-
minimum and maximum intensity is obtained from around 50 W/m2 to tallic frame. The layers and properties of materials of PV module are is
1100 W/m2. However, Malaysian weather is cloudy therefore data given in Table 3.

139
H. Fayaz et al. Solar Energy 179 (2019) 135–150

Fig. 3b. Cross-sectional view diagram of the thermal collector.

The investigation of PVT and PVT-PCM systems is conducted to 2.3.1. Thermophysical properties and characterisation of PCM
remove thermal energy from the PV module at different solar irradia- Phase change material used in this experiment is Paraffin with
tions from 200 W/m2 to 1000 W/m2 at the 32 °C set as the ambient commercial code A44-PCM. Thermophysical properties are determined
temperature throughout the experiment. Temperatures form centre using differential scanning calorimetry (DSC) during the melting cycle
points of glass, Tedlar, inlet and outlet of water for PV, PVT and PVT- as shown in Fig. 6. The material undergoes endothermic reaction at an
PCM are taken with thermocouples along with other data for the whole onset temperature of 44.77 °C which attributes to the change in phase
experiment. Changing solar irradiations effect on the cooling behaviour from solid to liquid form. During this phase change, a large amount of
of PV modules, electrical efficiency and maximum power output along energy has been absorbed by the paraffin wax as can be identified from
with thermal energy yield is investigated on PV, PVT, and PVT-PCM the heat capacity value. The change in Cp revealed the latent heat
systems by analysing achieved data. absorption characteristics of the paraffin wax. This Cp is responsible for
Table 4 shows the working parameters and weather conditions of energy storage in the material. Higher peak corresponds to high energy
the outdoor environment on average. Solar irradiance were recorded storage which in turn leads to the high performance of the system.
for the whole of the day and average solar irradiance values were ob- To further understand the stability of paraffin wax, TGA pattern has
tained from 200 W/m2 to 1000 W/m2 for an hour to investigate the been analyzed. Fig. 7 shows the TGA pattern of intended specimen. It
performance of the systems. can be observed that the material is stable till 200 °C. After 200 °C,

Fig. 4. The fabrication process of PVT and PVT-PCM systems.

140
H. Fayaz et al. Solar Energy 179 (2019) 135–150

Fig. 5. Outdoor experimental setup of PVT system.

Table 2 Table 4
Specifications of the PV module. Working outdoor parameters of the investigation at the location of the ex-
periment.
Item Specification
Working parameters Value Unit
Brand name EPV
Materials Polycrystalline silicon Irradiance 200–1000 W/m2
Area of a cell 0.02433 m2 Average relative humidity 75 %
Maximum power 305 W Average wind speed 0.53 m/s
Short circuit current (Isc) 8.92A Average ambient temperature 32 °C
Open circuit voltage (Voc) 45.7 V Average water inlet temperature 32 °C
Current at Pmax. (Impp) 8.17A Tilt angle 15 Degree
Voltage at Pmax. (Vmpp) 30.6 V
Operating temperature −40 °C - ± 85 °C
STC 1000 W/m2, AM 1.5, 25 °C
the other PCMs tested A44 paraffin wax was selected for the reasons
Tolerance 0∼+3%
that it’s safer and its life of degradation due to temperature range is
satisfactory, however, there is very little compromise on thermal con-
oxidation reactions cause weight loss of the specimen. Although spe- ductivity but that’s ignored as compared to other factors for the study.
cimen underwent a change in phase (melting) at 44.77 °C as evident in Moreover, Malaysian climatic conditions show that PV cell temperature
the DSC pattern, over the stability of the material has not been affected. can exceed from 25 °C to 70 °C+ in different weather conditions, which
No weight loss can be seen at 44.77 °C in TGA pattern. is harmful for the efficiency and life of PV modules. Therefore, to cut
A sharp weight loss took place at 230–250 °C. And the specimen the temperature rise at maximum temperature transition of phase
showed 100% weight loss at approximately 300 °C. This 100% weight change for PCM was chosen as 44 °C where the heat starts to be ab-
loss also indicates the purity level of the specimen under investigation. sorbed and stored by PCM (A44) at peak hot timings to keep cell
Thermophysical properties of PCM Paraffin (A44) are shown in Table 5. temperature of PV at minimum as much as possible. Under the desired
Five PCMs were inspected for the study, which are Decanoic Acid, other properties except thermal conductivity, the PCM (A44) is chosen
Lauric acid, Paraffin wax (A44), Paraffin wax-52 and I-Tetra Decnol. after its satisfactory results of temperature reduction of PV module in
The criterion of selection was safety in terms of toxicity, corrosiveness numerical analysis, which is the main aim of this study. There have
and other disastrous factors of PCM to be used in the project. Amongst been limitations for the other PCMs with all desired features designed

Table 3
PVT collector materials and thermal properties (Nahar et al., 2017a).
Layer Materials Heat capacity at constant pressure [J/(kg K)] Thermal Conductivity [W/(m K)] Density [kg/m3] Thickness (mm)

Top cover Glass 500 2 2450 3


Encapsulant EVA 2090 0.311 950 0.8
Solar cell Silicon 700 148 2329 0.1
Bottom cover Tedlar 1250 0.15 1200 0.05
Conductive medium Thermal paste 650 1.42 2500 0.2
Thermal collector Aluminium 2700 237 1
Working fluid Water 998 0.68 –

141
H. Fayaz et al. Solar Energy 179 (2019) 135–150

110 =N 1 (19)
100 PCM: A44 (Paraffin) Onset: 44.77 °C In this relationship, N represents the number of data. While, the
Heat Flow Endo Up (mW)

90 Peak: 45.73 precision is presented as;


Peak Height: 41.035 mW
80 Endset: 46.87 °C SF
R=
Heat of Fusion: 245.077 J/g N1/2 (20)
70
N
60 1
SF = (Xi X¯ ) 2
50 N i=1 (21)
40 In this equation, SF and (Xi X¯ ) 2 represent the standard deviation
30 and deviation of Xi . Further, the true value is presetned as;
0 10 20 30 40 50 60 70 80 90 100
Temperature(°C) X ' = X¯ ± U (95%) (22)

Fig. 6. Results of DSC tests of Paraffin A44. where X and X̄ are true temperature or flow and mean temperature or
'

flow measured.

2.5. Test procedure

The experimental test procedure is carried out in way that con-


sistent and stable data on average can be achieved for prescribed data
range and design parameters of the systems. The data was taken about
100 days and only refined data of desired values of working conditions
e.g. solar irradiations, wind speed etc. were taken for evaluation pur-
pose to compare and validate with numerical analysis. The days were
clear as well as cloudy and semi cloudy. Therefore, data on clear day for
desired amount of solar irradiation was taken on average. For example,
1000 W/m2 was achieved from 12 noon to 2 pm on average on sunny
clear days, further, 600 W/m2 was achieved from 10am to 12 pm or
3 pm to 5 pm on average at time period of clear sky. To make sure the
average amount of solar irradiation achieved with clouds interruptions,
the initial working conditions of the systems were taken into con-
sideration to reduce error percentage. Such work is done by some other
research as well as they used the similar procedure to analyze the data
(Nahar et al., 2017b; Rahman et al., 2017).

Fig. 7. TGA pattern of paraffin wax. 3. Results and discussion

Table 5 In this paper, numerical and experimental investigations on PV, PVT


Thermophysical properties of Paraffin (A44). and PVT-PCM collectors have been carried out to analyse the perfor-
Property Value mance under different operating conditions. A novel design of thermal
collector is introduced with the Aluminium material. A serpentine
Thermal conductivity 0.18 W/m K thermal collector design has been used as flow path of water passively
density 805 kg/m3
driven by overhead water tank head. The operating conditions of
Specific heat at constant pressure 2150 J/kg K
Temperature of transition 44 °C
200 W/m2 to 1000 W/m2 irradiation range is set in the present in-
Latent heat −242 kJ vestigation. The water inlet temperature and outdoor ambient tem-
Transitional interval 1 °C peratures are taken as 32 °C.

3.1. Optimization and selection of thermal collector diameter and material


for the research project as discussed earlier.
Fig. 8 shows the inner diameter effect on output temperature and
2.4. Uncertainty analysis thermal efficiency of the thermal collector when assembled with PV
under same working conditions of 1000 W/m2 and 0.5 LPM. Three
For the uncertainty analysis in this investigation, calibration and diameters such as 3.5, 7.5 and 15 mm are analysed to select for the
data acquisition errors are taken into account in evaluating overall further numerical and experimental investigation. The effect of each
measurement error. In addition, to evaluate the data reduction error, diameter is presented in terms of output temperature and thermal ef-
which is made up of thermocouple error and bias error of flow meter, ficiency separately. It is observed that diameter has very significant
the later one is not considered for the reason that the value is very is effect on the performance of thermal collector in heat transfer effi-
very small and can be negligible. The confidence interval of 95% is ciency. The rend shows the increasing thermal performance with re-
considered for uncertainty analysis as shown in following Eq. (18). spect to decreasing inner diameter of the thermal collector (Joshi and
B 2 + (t Khandwawala, 2014).
U= ,95% × R) (18)
Therefore, from the results, it can be concluded that lower the
where U, B, and tλ,95% R are measurement uncertainty, bias error, and diameter the higher is the efficiency of the thermal collector. Such a
an estimate of the precision error in the repeated temperature and flow significant difference in thermal performance of the thermal collector
measurement at 95% confidence. Where the degree of freedom is given due to working diameter pushes to be careful in selecting the diameter
as λ represented below as; of the thermal collectors. Conversely, it is not necessary that lowest

142
H. Fayaz et al. Solar Energy 179 (2019) 135–150

similar. Furthermore, the cost of copper is higher than aluminium due


to high demand and shortage in the market. The output temperature
and thermal efficiency of aluminium and copper thermal collectors are
observed as 61.11 & 62 °C, and 69.1 & 71.24% respectively (Fayaz
et al., 2018). Therefore, keeping all factors for selection, aluminium is
selected for further numerical and experimental investigation.

3.2. Solar irradiation reception potential profile

Whole day solar irradiation incident on the collectors used for ex-
periments are shown in Fig. 10. The significant angle and intensity of
solar irradiations on average are achieved from 08:30 approximately,
where PV panel starts to generate power. Furthermore, PV panels kept
producing power till 18:45 approximately when the angle and intensity
of solar irradiation were so acute to provide a significant amount of
solar irradiation to PV panels for power production. It can further be
seen that sky in Malaysia is not 100% clear due to clouding because the
solar irradiation suddenly drops at some time during the whole day. It is
Fig. 8. Optimization and selection of thermal collector diameter. rare to get clear sky in Malaysian climatic conditions. However, there is
the great potential of solar irradiations availability in Malaysia as can
be seen in Fig. 10.
diameter should be selected for the better thermal performance of
The maximum solar irradiations are achieved as 1050 W/m2 ap-
thermal collector because the lowest diameter has other compromises,
proximately for the peak of the day starting from 12:00 to about 14:00.
e.g. limiting mass flow rates, chances of increased system pressure and
This period of solar irradiation is very ideal for thermal as well as
rapid blockage due to possible fouling and scaling inside the thermal
electrical energy gain. However, on average, the best timing for the
collector. Thus, the selection of the inner diameter in this investigation
electrical as well as thermal energy gain starts from 10:30 to 16:30
is limited to 7.5 mm. The reason behind this selection is to avoid the
where solar irradiations are achieved as 600 W/m2 or above on a clear
exceeding pressure in the system and possible blockage due to fouling
day.
and scaling, which may lead the system to failure and major main-
tenance costs. Higher diameter is not selected due to the reason that it
3.3. Effect of solar irradiation on PV, PVT and PVT-PCM systems’
has considerably low thermal performance as compared to 7.5 mm and
temperature profile
very high-mass flow rates are not required in the domestic hot-water
system due to a certain limit of hot-water consumption with preferred
The different solar irradiations intensities have specifically a major
higher output temperatures (Fayaz et al., 2018).
varying effect in heating the whole PV, PVT and PVT-PCM systems
Fig. 9 shows the material optimization and selection for the current
along with other influencing working parameters. In Fig. 11(a)–(e), the
investigation under same working conditions. Aluminium and copper
surface temperature flow over the PV is presented for its exposure to
are analyzed and compared for their thermal performance to go with
solar irradiations of from 200 W/m2 to 1000 W/m2 at natural convec-
suitable and desired one. The copper has better thermal conductivity as
tion conditions. From the 200 to 1000 W/m2 solar irradiations, the
compared to the aluminium material (Hust and Lankford, 1984). It is
highest values of PV temperatures are found as approximately 77.9 °C
seen in Fig. 9 that there is the difference in the aluminium and copper in
and 37 °C respectively with the surrounding environmental temperature
their heat transfer efficiency but not as big as to comprise with cost,
of 32 °C. Additionally, the edges of the PV module are observed cooler
weight, workability and availability of each of them. Consequently, for
than the middle part of the PV, which is fewer degrees Celsius lower in
the selection of copper material for some heat transfer efficiency, there
value. The reason behind the cooler edges of PV is three directional heat
is a compromise on the other factors. Aluminium is lightweight, which
removals to the surroundings as compared to the middle part which
is almost three times lighter than copper, which reduces the overall
only dissipated heat from glass and Tedlar only. The electrical effi-
weight of the PVT system if the length of the thermal absorber is kept
ciency of the PV module is found as 11.44 and 11.61% for experimental
and numerical respectively at 1000 W/m2. However, at reducing irra-
diations electrical efficiency is improved due to lower cell temperatures
associated with solar irradiations intensity (Jaaz et al., 2018).
Similarly, in Figs. 12(a)-(j) and 13(a)-(j), the effect of solar irra-
diation on the PVT and PVT-PCM systems’ surface temperature and the
temperature of water flowing and the pipe is shown in 3D simulation
plots. The temperature distribution is depicted for the solar irradiations
ranging from 200 W/m2 to 1000 W/m2 at the fixed average water inlet
temperature of the water as 32 °C and 0.5 LPM mass flow rate. It can be
seen that temperature of the PVT system decreased approximately from
75.1 °C to 45.9 °C at decreasing solar irradiation from 200 W/m2 to
1000 W/m2 (Nahar et al., 2017b). Furthermore, the temperature falls
steadily with a considerable difference as the irradiation intensity de-
creases. The temperature distribution of heat transferring fluid flowing
in the pipe at varying solar irradiations from 200 W/m2 to 1000 W/m2
at fixed 0.5 LPM is shown in Fig. 12(f)-(j). The water channel captures
heat from 32 °C to 63.1 °C at 1000 W/m2. At these irradiation ranges the
module temperature is reduced from 75 °C to 46 °C. There is an almost a
uniform drop in the temperature PV as the irradiation levels decrease.
Fig. 9. Optimization and selection of thermal collector material. Similarly, for PVT-PCM in Fig. 13(a)-(j), the temperature drops in

143
H. Fayaz et al. Solar Energy 179 (2019) 135–150

Fig. 10. Incident solar irradiation on a typical day.

the similar pattern as compared to PVT with respect to varying solar 100 W/m2 is obtained as 2.32 °C and 2.3 °C for experimental and nu-
irradiations. However, in PVT-PCM, the temperature decreased more merical cases respectively. Maximum cell temperature reduction of
than the PVT because the phase change material absorbs a certain 8.3 °C and 8.1 °C in case of PVT, and 12.8 °C and 12 °C is achieved in
amount of heat and stores it. Therefore, some heat is carried out with case of the PVT-PCM system from the PV module.
working fluid which is water and some heat is stored in phase change The incident irradiation effect on power and electrical efficiency is
material. Thus, it causes PVT-PCM to reduce the PV temperatures more depicted in Figs. 15a and 15b. It can be seen that with the increase in
as compared to PVT. The maximum temperature of the system is shown irradiations there is a corresponding increase in output power due to
as 70.7 °C at 1000 W/m2, and minimum temperature is accumulated as the reason that both current and voltage increase, however current
45.3 °C at the solar irradiations of 200 W/m2. Furthermore, in increases too higher than voltage rate of increment. Thus, the voltage
Fig. 13(f)-(j), the streamline temperature is shown for the water flowing shows a linear increment with increasing irradiation intensities
channel. At 1000 W/m2, the achieved temperature is 59.4 °C and a (Başoğlu and Çakır, 2016; Radziemska, 2003). In Fig. 15a output power
minimum temperature of 41 °C is obtained at 200 W/m2. The output of PV is significantly lower than PVT system, which is due to the reason
temperature from the PVT-PCM is smaller than the PVT for the reason of higher cell temperature accumulation in the PV module due to its
of storing capacity of phase change materials used in this experiment. lower rate of heat removal from its cell temperature in natural con-
vection only. Conversely, PVT system acquires higher output power
3.4. Analysis of electrical performance of PV, PVT and PVT-PCM at because of the reason of its cooling, which removes the excessive heat
varying irradiation levels. from cell temperature hence creating higher power output.
PVT-PCM removes maximum heat as compared to PVT, which re-
Fig. 14 shows the effect of the solar irradiation on the cell tem- moves the excessive heat from cell temperature as PCM stores heat
perature. The average cell temperature increases with increasing irra- along with water cooling, hence creating higher power output.
diation intensity for PV, PVT and PVT-PCM numerically and experi- In the PV, PVT and PVT-PCM cases, numerical values are almost
mentally. As the incident irradiation increases the excessive heat stable, however; experimental deviations are noted significantly at
generates in the PV module causing the increase in the average tem- lower and higher irradiation level.
perature of PV cell (Nordmann and Clavadetscher, 2003). It can be As the solar irradiation increases the electrical efficiency decreases
observed from the Fig. 14 that there are some deviations in an ex- because of the reason that cell temperature increases prominently at
perimental curve as compared to numerical curve because of the ex- high incoming radiation flux as shown in Fig. 15b. For PV, electrical
perimental control parameters errors in outdoor weather conditions efficiency decreases from 13.72% to 11.62 in the numerical case and
and sensors uncertainties. The working conditions for outdoor data are 13.56 to 11.44% in the experimental case from 200 W/m2 to 1000 W/
measured at 32 °C for ambient and inlet water temperature at a fixed m2 irradiation intensities. In the like manner for PVT, values of elec-
mass flow rate of 0.5LPM. It can be seen that the PVT-PCM system trical efficiency decrease from 13.85% to 12.16% for the numerical case
achieves best results with minimum cell temperature gain. It is because and 13.74% to 12% for the experimental case. For PVT-PCM, values of
of the reason that PCM (A44) absorbs a certain amount of heat latently electrical efficiency decrease from 14% to 12.43% for the numerical
(Browne et al., 2016), which reduces the temperature of a cell tem- case and 13.87% to 12.31% for the experimental case. Further, at every
perature of the PVT-PCM system. Cell temperature achieved by PV is 100 W/m2 interval PV, electrical efficiency is decreased at the average
observed as a maximum of 77.6 °C and 75.1 °C for experimental and rate of 0.21% and 0.212% for numerical and experimental cases re-
numerical cases, and minimum 46.3 °C and 44 °C for experimental and spectively. Moreover, for PVT, decrement in electrical efficiency is
numerical cases respectively. However, cell temperature of PVT is much found as 0.169% and 0.174% numerically and experimentally respec-
lower as compared to PV module, which is observed as highest 69.3 °C tively.
and 67 °C, and lowest 43.6 °C and 42 °C for experimental and numerical
cases respectively. Additionally, an increment in cell temperature on 3.5. Thermal performance of PVT and PVT-PCM system at different
average at each 100 W/m2 is obtained as 2.6 °C and 2.5 °C experimen- irradiation intensities
tally and numerically respectively. Furthermore, the cell temperature of
PVT-PCM is observed as highest 64.8 °C and 63 °C, and the lowest The solar irradiation effect on the thermal performance of the PVT
41.6 °C and 40 °C for experimental and numerical cases respectively. system is shown in Figs. 16a, 16b, 17a and 17b. The water outlet
Additionally, an increment in cell temperature on average at each temperature increases with increase in irradiation level from 200 W/m2

144
H. Fayaz et al. Solar Energy 179 (2019) 135–150

(a) G= 1000 W/m2 (b) G= 800 W/m2

(c) G= 600 W/m2 (d) G= 400 W/m2

(e) G= 200 W/m2


Fig. 11. PV module temperature distribution for the effect of irradiation at 200 to1000 W/m2.

to 1000 W/m2 numerically and experimentally as seen in Fig. 16a. Such the systems and the environment. Therefore, some of the heat is rapidly
increasing trend is ascribed to the enhancing conduction and convective dissipated into the environment due to free convection phenomenon. It
heat transfer rate in between PV layers and from Tedlar to aluminium gets lesser time for transferring heat efficiently to working fluid due to
heat exchanger hence finally to flowing water due to the higher tem- high temperatures and lower ambient temperature conditions of 32 °C
perature gradient (Rohsenow et al., 1998). Higher the temperature (Kreith and Black, 1980). It can be observed from the same figure that
gradient allows the higher heat transfer rate. Therefore, as irradiation there is a good accord in numerical and experimental results and con-
increases the more energy is transmitted to working fluid at higher tinues with the similar trend from some deviation after 400 W/m2 in
rates. In result, higher output temperatures are acquired. The rate of case of experimental data. Output temperature values for PVT system
heat transfer is uniform from 200 W/m2 to 1000 W/m2 for numerical are obtained as 61.11 °C at 1000 W/m2 and 38.92 °C at 200 W/m2 for
values but it changes in curved shape in declining position after 400 W/ the numerical case, whereas, 59.88 °C at 1000 W/m2 and 38.85 °C at
m2 for the experimental case. It is because of the reason that at higher 200 W/m2 for the experimental case. Output temperature increment at
irradiation levels PVT and PVT-PCM systems accumulated a high every 100 W/m2 is achieved as 2.219 °C and 2.103 °C on average for
amount of heat, which created a high-temperature gradient between numerical and experimental cases respectively. For PVT-PCM, output

145
H. Fayaz et al. Solar Energy 179 (2019) 135–150

(f) G= 10000 W/m2 (a) G= 1000 W/m2

(g) G= 800 W/m2 (b) G= 800 W/m2

(h) G= 600 W/m2 (c) G= 600 W/m2

(i) G= 400 W/m2 (d) G= 400 W/m2

(j) G= 200 W/m2 (e) G= 200 W/m2

Fig. 12. Surface and streamline temperature distribution over through PVT system.

temperature values are obtained as 59.76 °C at 1000 W/m2 and 38.14 °C average numerically and experimentally respectively. Thermal energy
at 200 W/m2 for the numerical case, whereas, 58.1 °C at 1000 W/m2 relationship on solar irradiation is shown in Fig. 16b at the solar irra-
and 38 °C at 200 W/m2 for the experimental case. Output temperature diation range from 200 W/m2 to 1000 W/m2. At the high-temperature
increment at every 100 W/m2 is achieved as 2.16 °C and 2.06 °C on gradient caused by high solar irradiation, the heat transfer between the

146
H. Fayaz et al. Solar Energy 179 (2019) 135–150

(f) G= 200 W/m2 (a) G= 200 W/m2

(g) G= 200 W/m2 (b) G= 200 W/m2

(h) G= 200 W/m2 (c) G= 200 W/m2

(i) G= 200 W/m2 (d) G= 200 W/m2

(j) G= 200 W/m2 (e) G= 200 W/m2

Fig. 13. Surface and streamline temperature distribution over through PVT-PCM system.

thermal collector and water increases considerably. Therefore, as obtained as 999.6 W and 928.2 W at 1000 W/m2 for numerically and
compared to 200 W/m2 more thermal energy is obtained at 1000 W/m2 experimentally respectively. In addition, smallest thermal energy va-
due to increased convective and conductive heat transfer within the lues are achieved as 142.8 W and 135.66 W at 200 W/m2 for numeri-
system. The maximum value of thermal energy for PVT systems is cally and experimentally respectively. For each 100 W/m2 irradiations,

147
H. Fayaz et al. Solar Energy 179 (2019) 135–150

Fig. 14. Effect of irradiation on cell temperature of PV, PVT and PVT-PCM.

Fig. 16a. Effect of solar irradiation on the output temperature of PVT and PVT-
PCM.

Fig. 15a. Effect of solar irradiation on the output power of PV, PVT and PVT-
PCM.

Fig. 16b. Effect of solar irradiation on thermal energy of PVT and PVT-PCM.

Fig. 15b. Effect of solar irradiation on the electrical efficiency of PV, PVT and
Fig. 17a. Effect of solar irradiation on the thermal efficiency of PVT and PVT-
PVT-PCM.
PCM.

148
H. Fayaz et al. Solar Energy 179 (2019) 135–150

200 W/m2 for PVT system. Further, lowest thermal efficiency is ob-
tained as 81.7% and 78.46% for numerically and experimentally re-
spectively. In the same manner, the highest values for overall efficiency
numerically and experimentally are 87.43% and 85.53% respectively
for the PVT-PCM system, at 200 W/m2. Finally, the lowest overall ef-
ficiency obtained is 78.89% and 76.53% numerically and experimen-
tally respectively at 1000 W/m2. Numerical and experimental results
are observed with no a big difference or deviation and are in the sa-
tisfactory range of mathematical 3D model proposed for this in-
vestigation.

4. Conclusion

In this investigation, evaluation of the numerical and experimental


performance of a PV, PV/T and PVT-PCM systems with the serpentine
design of aluminium pipe is presented. Numerical analysis based on
FEM is done by using COMSOL Multiphysics® along with outdoor ex-
periments conducted under controlled operating conditions. The main
findings achieved from the present research work are given below:

Fig. 17b. Effect of solar irradiation on the overall efficiency of PVT and PVT-
PCM. • For the confirmation of the thermal and electrical efficiency of the
proposed design, an experimental verification is carried out. The
tests show that the experimental results are in satisfactory range
thermal energy on average is achieved as 85.68 W and 79.254 W for with the numerical analysis for different mass flow rates.
numerically and experimentally respectively. Similarly, for PVT-PCM
maximum thermal energy is gained as 985.67 W and 860.37 W at
• For PV, the maximum achieved cell temperature is 75.6 and 75.1 °C
PVT, and for PVT, it is 67 and 69.3 °C numerically and experimen-
1000 W/m2 for numerically and experimentally respectively. In addi- tally. In addition, PVT-PCM, maximum cell temperature is achieved
tion, smallest values for thermal energy are achieved as 216.69 W and as 63 and 64.8 °C numerically and experimentally respectively.
170.64 W at 200 W/m2 for numerically and experimentally respec-
tively. At the same interval of 100 W/m2, thermal energy is achieved as
• Maximum cell temperature reduction of 8.3 °C and 8.1 °C in case of
PVT, and 12.8 °C and 12 °C is achieved in the case of the PVT-PCM
76.898 W and 68.973 W on average for numerically and experimentally system from the PV module.
respectively.
Fig. 17a shows that thermal efficiency at the irradiation level of
• The electrical efficiency of PV, PVT and PVT-PCM numerically is
achieved 13.72, 13.85 and 14% and for the experimental in-
200 W/m2 to 1000 W/m2. Opposite to thermal energy graphs, thermal vestigation; it is obtained as 13.56%, 13.74% and 13.87% at 200 W/
energy is decreased with an increase in solar irradiation. That’s be- m2 respectively.
cause, at the higher irradiation levels, heat obtained by the system is
not efficiently transferred to the water as some of the heat is lost in the
• For PVT, the highest overall efficiency is 89.9 and 88.84%, and for
PVT-PCM, the maximum overall efficiency is obtained as 85.88 and
convection process to the surroundings. The heat convection phenom- 82.87% numerically and experimentally respectively.
enon happens so fast that it provides less time for heat to transfer from
PV layers to the thermal collector and hence finally to the water flowing
• In the case of the PVT system, electrical performance is improved as
6.2 and 4.8% and in the case of PVT-PCM; it is improved as 7.2 and
inside the collector. Temperature difference which is responsible for 7.6% numerically and experimentally respectively.
excessive heat transfer in between systems layers to the water is also
responsible for heat flow to the surrounding due to a temperature
• The highest value for the electrical efficiency of PVT is achieved as
13.744 and 113.852%, and similarly, for PVT-PCM, electrical effi-
gradient. Conversely, at lower irradiation levels, thermal efficiency is ciency is achieved as 13.829 and 13.987% at 200 W/m2 numerically
improved and is higher than irradiations at high levels. and experimentally respectively.
The maximum thermal efficiency achieved for the PVT system is as
76.1% and 75.1% at 200 W/m2 numerically and experimentally re- Acknowledgement
spectively. Additionally, at 1000 W/m2, minimum values of thermal
energy are achieved as 69.1% and 66.2% numerically and experimen- The authors would like to acknowledge the financial support from
tally respectively. Thermal efficiency drop is observed as 0.7% and Higher Institution Centre of Excellence (HICoE) Grant, Ministry of
0.9% at each interval of 100 W/m2 numerically and experimentally Higher Education (Project: UM.0000067/HME.OM, UMPEDAC - 2016)
respectively. Similarly, for the PVT-PCM, thermal efficiency at highest to carry out this research.
value is achieved as 72.83% and 71.1% at 200 W/m2 numerically and
experimentally respectively. Additionally, at 1000 W/m2 minimum References
thermal efficiency is achieved as 65.91% and 63.25% numerically and
experimentally respectively. Drop in thermal efficiency is achieved as Aberoumand, Sadegh, Ghamari, Shahin, Shabani, Bahman, 2018. Energy and exergy
0.69% and 0.78% at every 100 W/m2 numerically and experimentally analysis of a photovoltaic thermal (PV/T) system using nanofluids: An experimental
respectively. study. Sol. Energy 165, 167–177.
Adiyabat, Amarbayar, Kurokawa, Kosuke, Otani, Kenji, Enebish, Namjil, Batsukh,
Fig. 17b shows that the overall efficiency of the systems with the Garmaa, Battushig, Mishiglunden, ... Ganbat, Bathuu, 2006. Evaluation of solar en-
effect of varying solar irradiations. Overall efficiency shows the similar ergy potential and PV module performance in the Gobi Desert of Mongolia. Prog.
trend as compared to electrical and thermal efficiencies of the systems. Photovoltaics Res. Appl. 14 (6), 553–566.
Ahmed, Ferdous, Al Amin, Abul Quasem, Hasanuzzaman, M., Saidur, R., 2013.
Overall efficiency is the addition of electrical and thermal efficiencies Alternative energy resources in Bangladesh and future prospect. Renew. Sustain.
where the major part is the thermal efficiency, hence dominating the Energy Rev. 25, 698–707.
similar trend of fall in overall efficiency with respect to increasing solar Al-Waeli, A.H.A., Chaichan, M.T., Sopian, K., Kazem, H.A., 2018a. Comparison study of
indoor/outdoor experiments of SiC nanofluid as a base-fluid for a photovoltaic
irradiations. Overall efficiency at the maximum values is achieved as thermal PV/T system enhancement. Energy 151, 33–44.
90.08% and 88.95% numerically and experimentally respectively at Al-Waeli, Ali H., Sopian, K., Kazem, Hussein A., Chaichan, Miqdam T., 2016. Photovoltaic

149
H. Fayaz et al. Solar Energy 179 (2019) 135–150

solar thermal (PV/T) collectors past, present and future: A. Int. J. Appl. Eng. Res. 11 Kreith, Frank, & Black, William Z., 1980. Basic heat transfer: Harper & Row New York.
(22), 10757–10765. Lim, Jong-Log, Woo, Sung-Cheol, Jung, Tae-Hee, Min, Yong-Ki, Won, Chang-Sub, Ahn,
Al-Waeli, Ali H.A., Sopian, K., Chaichan, Miqdam T., Kazem, Hussein A., Ibrahim, Adnan, Hyung-Keun, 2013. Analysis of factor on the temperature effect on the output of PV
Mat, Sohif, Ruslan, Mohd Hafidz, 2017. Evaluation of the nanofluid and nano-PCM module. The Transactions of The Korean Institute of Electrical Engineers 62 (3),
based photovoltaic thermal (PVT) system: An experimental study. Energy Convers. 365–370.
Manage. 151, 693–708. https://doi.org/10.1016/j.enconman.2017.09.032. Malvi, C.S., Dixon-Hardy, D.W., Crook, R., 2011. Energy balance model of combined
Al-Waeli, Ali H.A., Sopian, K., Kazem, Hussein A., Chaichan, Miqdam T., 2018b. photovoltaic solar-thermal system incorporating phase change material. Sol. Energy
Nanofluid based grid connected PV/T systems in Malaysia: A techno-economical 85 (7), 1440–1446.
assessment. Sustain. Energy Technol. Assess. 28, 81–95. Mousavi, Soroush, Kasaeian, Alibakhsh, Shafii, Mohammad Behshad, Jahangir,
Al-Waeli, Ali H.A., Sopian, K., Kazem, Hussein A., Yousif, Jabar H., Chaichan, Miqdam T., Mohammad Hossein, 2018. Numerical investigation of the effects of a copper foam
Ibrahim, Adnan, Ruslan, Mohd Hafidz, 2018c. Comparison of prediction methods of filled with phase change materials in a water-cooled photovoltaic/thermal system.
PV/T nanofluid and nano-PCM system using a measured dataset and Artificial Neural Energy Convers. Manage. 163, 187–195.
Network. Sol. Energy 162, 378–396. Nahar, Afroza, Hasanuzzaman, M., Rahim, N.A., 2017a. Numerical and experimental
Başoğlu, Mustafa Engin, Çakır, Bekir, 2016. Comparisons of MPPT performances of iso- investigation on the performance of a photovoltaic thermal collector with parallel
lated and non-isolated DC–DC converters by using a new approach. Renew. Sustain. plate flow channel under different operating conditions in Malaysia. Sol. Energy 144
Energy Rev. 60, 1100–1113. (Supplement C), 517–528. https://doi.org/10.1016/j.solener.2017.01.041.
Bertram, Erik, Glembin, Jens, Rockendorf, Gunter, 2012. Unglazed PVT collectors as Nahar, Afroza, Hasanuzzaman, M., Rahim, N.A., 2017b. Numerical and experimental
additional heat source in heat pump systems with borehole heat exchanger. Energy investigation on the performance of a photovoltaic thermal collector with parallel
Procedia 30, 414–423. https://doi.org/10.1016/j.egypro.2012.11.049. plate flow channel under different operating conditions in Malaysia. Sol. Energy 144,
Browne, Maria C., Lawlor, Keith, Kelly, Adam, Norton, Brian, Cormack, Sarah J.M., 2015. 517–528. https://doi.org/10.1016/j.solener.2017.01.041.
Indoor characterisation of a photovoltaic/thermal phase change material system. Nasrin, R., Hasanuzzaman, M., Rahim, N.A., 2018a. Effect of high irradiation and cooling
Energy Procedia 70 (Supplement C), 163–171. https://doi.org/10.1016/j.egypro. on power, energy and performance of a PVT system. Renew. Energy 116 (Part A),
2015.02.112. 552–569. https://doi.org/10.1016/j.renene.2017.10.004.
Browne, Maria C., Quigley, Declan, Hard, Hanna R., Gilligan, Sarah, Ribeiro, Nadja C.C., Nasrin, R., Rahim, N.A., Fayaz, H., Hasanuzzaman, M., 2018b. Water/MWCNT nanofluid
Almeida, Nicholas, McCormack, Sarah J., 2016. Assessing the thermal performance of based cooling system of PVT: Experimental and numerical research. Renew. Energy
phase change material in a photovoltaic/thermal system. Energy Procedia 91, 121, 286–300. https://doi.org/10.1016/j.renene.2018.01.014.
113–121. https://doi.org/10.1016/j.egypro.2016.06.184. Nishioka, Kensuke, Hatayama, Tomoaki, Uraoka, Yukiharu, Fuyuki, Takashi, Hagihara,
Chandrasekar, M., Suresh, S., Senthilkumar, T., 2013. Passive cooling of standalone flat Ryuzou, Watanabe, Momoki, 2003. Field-test analysis of PV system output char-
PV module with cotton wick structures. Energy Convers. Manage. 71, 43–50. acteristics focusing on module temperature. Sol. Energy Mater. Sol. Cells 75 (3),
Chauhan, Aditya, Tyagi, V.V., Anand, Sanjeev, 2018. Futuristic approach for thermal 665–671.
management in solar PV/thermal systems with possible applications. Energy Convers. Nordmann, Thomas, Clavadetscher, Luzi, 2003. Understanding temperature effects on PV
Manage. 163, 314–354. system performance. Paper Presented at the Photovoltaic Energy Conversion, 2003.
Daghigh, Ronak, Ibrahim, Adnan, Jin, Goh Li, Ruslan, Mohd Hafidz, Sopian, Proceedings of 3rd World Conference on.
Kamaruzzaman, 2011. Predicting the performance of amorphous and crystalline si- Nouira, Meriem, Sammouda, Habib, 2018. Numerical study of an inclined photovoltaic
licon based photovoltaic solar thermal collectors. Energy Convers. Manage. 52 (3), system coupled with phase change material under various operating conditions. Appl.
1741–1747. Therm. Eng. 141, 958–975.
Dupeyrat, P., Ménézo, C., Fortuin, S., 2014. Study of the thermal and electrical perfor- Radziemska, E., 2003. The effect of temperature on the power drop in crystalline silicon
mances of PVT solar hot water system. Energy Build. 68 (Part C(0)), 751–755. solar cells. Renew. Energy 28 (1), 1–12.
https://doi.org/10.1016/j.enbuild.2012.09.032. Rahman, M.M., Hasanuzzaman, M., Rahim, N.A., 2015. Effects of various parameters on
Ebadi, Soroush, Tasnim, Syeda Humaira, Aliabadi, Amir A, Mahmud, Shohel, 2018. PV-module power and efficiency. Energy Convers. Manage. 103 (Supplement C),
Melting of nano-PCM inside a cylindrical thermal energy storage system: Numerical 348–358. https://doi.org/10.1016/j.enconman.2015.06.067.
study with experimental verification. Energy Convers. Manage. 166, 241–259. Rahman, Mohammad Mafizur, Hasanuzzaman, Md., Rahim, Nasrudin Abd, 2017. Effects
Fayaz, H., Rahim, N.A., Saidur, R., Solangi, K.H., Niaz, H., Hossain, M.S., 2011. Solar of operational conditions on the energy efficiency of photovoltaic modules operating
energy policy: Malaysia vs developed countries. In: Paper Presented at the Clean in Malaysia. J. Cleaner Prod. 143, 912–924. https://doi.org/10.1016/j.jclepro.2016.
Energy and Technology (CET), 2011 IEEE First Conference on. 12.029.
Fayaz, H., Nasrin, R., Rahim, N.A., Hasanuzzaman, M., 2018. Energy and exergy analysis Riffat, Saffa B., Cuce, Erdem, 2011. A review on hybrid photovoltaic/thermal collectors
of the PVT system: Effect of nanofluid flow rate. Sol. Energy 169, 217–230. https:// and systems. Int. J. Low-Carbon Technol. 6 (3), 212–241.
doi.org/10.1016/j.solener.2018.05.004. Rohsenow, Warren M., Hartnett, James P., Cho, Young I., 1998. Handbook of Heat
Hasanuzzaman, M., Al-Amin, Abul Quasem, Khanam, Shamsunnahar, Hosenuzzaman, M., Transfer. McGraw-Hill, New York.
2015. Photovoltaic power generation and its economic and environmental future in Rostami, Zakie, Rahimi, Masoud, Azimi, Neda, 2018. Using high-frequency ultrasound
Bangladesh. J. Renewable Sustain. Energy 7 (1), 013108. waves and nanofluid for increasing the efficiency and cooling performance of a PV
Hottel, Hoyte, & Whillier, Austin, 1955. Evaluation of flat-plate solar collector perfor- module. Energy Convers. Manage. 160, 141–149.
mance. In: Paper Presented at the Trans. Conf. Use of Solar Energy. Skoplaki, Elisa, Palyvos, John A., 2009. On the temperature dependence of photovoltaic
Hust, Jerome G., Lankford, Alan B., 1984. Thermal conductivity of aluminum, copper, module electrical performance: A review of efficiency/power correlations. Sol.
iron, and tungsten for temperatures from 1 K to the melting point: National Bureau of Energy 83 (5), 614–624.
Standards, Boulder, CO (USA). Chemical Engineering Science Div. Song, Mengjie, Niu, Fuxin, Mao, Ning, Hu, Yanxin, Deng, Shiming, 2018. Review on
Jaaz, Ahed Hameed, Sopian, Kamaruzzaman, Gaaz, Tayser Sumer, 2018. Study of the building energy performance improvement using phase change materials. Energy
electrical and thermal performances of photovoltaic thermal collector-compound Build. 158 (Supplement C), 776–793. https://doi.org/10.1016/j.enbuild.2017.10.
parabolic concentrated. Results Phys. 9, 500–510. https://doi.org/10.1016/j.rinp. 066.
2018.03.004. Vokas, G., Christandonis, N., Skittides, F., 2006. Hybrid photovoltaic–thermal systems for
Joshi, Rajesh, Khandwawala, A.I., 2014. A comparative study of the effect of variation of domestic heating and cooling—A theoretical approach. Sol. Energy 80 (5), 607–615.
inside diameter of condenser and mass flow rate on the heat transfer coefficient in a Wolf, Martin, 1976. Performance analyses of combined heating and photovoltaic power
domestic refrigerator. J. Eng. Res. Appl. 4 (2), 514–518. systems for residences. Energy Convers. 16 (1–2), 79–90.

150

You might also like