Accepted Manuscript Effect of high irradiation and cooling on power, energy and performance of a pvt system
R. Nasrin, M. Hasanuzzaman, N.A. Rahim PII:
S0960-1481(17)30965-5
DOI:
10.1016/j.renene.2017.10.004
Reference:
RENE 9292
To appear in:
Renewable Energy
Received Date:
01 June 2017
Revised Date:
25 September 2017
Accepted Date:
01 October 2017
Please cite this article as: R. Nasrin, M. Hasanuzzaman, N.A. Rahim, Effect of high irradiation and cooling on power, energy and performance of a pvt system, Renewable Energy (2017), doi: 10.1016 /j.renene.2017.10.004
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT
Highlights: 1. Concentrating solar irradiation has been extrapolated from cell to module level. 2. A new correlation is developed for cell temperature calculation. 3. Every 100 W/m2 increase in irradiation raises electrical power by 19.65 W. 4. Optimum cooling rate suitable for radiations as high as 5000 W/m2 is 180 L/h. 5. Every 10 L/h increase in coolant flow rate lower cell temperature by 1.1oC.
ACCEPTED MANUSCRIPT 1 2
EFFECT OF HIGH IRRADIATION AND COOLING ON POWER, ENERGY AND PERFORMANCE OF A PVT SYSTEM
3
R. Nasrin,a,b, M. Hasanuzzamana, N.A. Rahima,c aUM
4 5 6 7 8 9 10 11
Power Energy Dedicated Advanced Centre (UMPEDAC), Level 4, Wisma R&D, University of Malaya, 59990 Kuala Lumpur, Malaysia bDepartment of Mathematics, Bangladesh University of Engineering and Technology, Dhaka-1000, Bangladesh c Renewable Energy Research Group, King Abdulaziz University, Jeddah 21589, Saudi Arabia Abstract
12
Irradiation level is the key factor of photovoltaic power generation. Photovoltaic/thermal
13
systems are more effective at concentrating power in areas of high irradiation as compared to
14
traditional PV systems. High irradiation maintains the cell temperature and maximizes
15
electrical-thermal energy. An optimum cooling system is required to remove the extra heat
16
from a PVT system, leading to enhancement of overall performance. In this research, the
17
effect of different high irradiation levels and cooling fluid flow rate are investigated in terms
18
of cell temperature, outlet temperature, electrical-thermal energy and overall performance of
19
PVT system. Finite element based software COMSOL Multiphysics has been used to solve
20
the problem numerically in three-dimensional model. The numerical model has been
21
validated with available experimental and numerical results. It is found that overall efficiency
22
increases with increasing fluid flow rate and with an optimum cooling fluid flow rate of about
23
180 L/h. Electrical and thermal energy increase from 197 to 983 W and 1165 to 5387 W
24
respectively, for increasing irradiation from 1000 to 5000 W/m2 with an optimized flow rate
25
of 180 L/h. Electrical, thermal and overall efficiency are found to be about 10.6, 71 and
26
81.6% respectively, at the highest irradiation level of 5000 W/m2.
27
Keywords: PVT system; High irradiation; Cooling; Power; Energy; Performance.
28
Nomenclature
29
A
Area of PVT surface (m2)
30
Al
Area of each lens (m2)
31
Asc
Area of each solar cell (m2)
32
Cp
Specific heat at constant pressure (Jkg-1K-1)
33
Cr
Concentration ratio
Corresponding author. Email address:
[email protected] (Dr. Rehena Nasrin)
1
ACCEPTED MANUSCRIPT 34
Ep
Electrical power (W)
35
Er
Received energy by PV (W)
36
Et
Thermal energy (W)
37
G
Solar radiation (Wm-2)
38
k
Thermal conductivity (Wm-1K-1)
39
m
Mass flow rate (kgs-1)
40
PV
Photovoltaic
41
PVT
Photovoltaic thermal
42
Psc
Packing factor of module
43
Tamb
Temperature of ambient (°C)
44
Tf
Temperature of water (°C)
45
Tg
Temperature of glass (°C)
46
The
Temperature of heat exchanger (°C)
47
Tin
Temperature of input water (°C)
48
Tout
Temperature of output water (°C)
49
Tr
Reference temperature (°C)
50
Tsc
Temperature of solar cell (°C)
51
Ttd
Temperature of tedlar (°C)
52
Uhea
Heat transfer coefficient from heat exchanger to ambient (Wm-2K-1)
53
Uhe
Heat transfer coefficient from heat exchanger to water (Wm-2K-1)
54
Uga
Heat transfer coefficient from glass to ambient (Wm-2K-1)
55
Ut
Heat transfer coefficient inside PV layers (Wm-2K-1)
56
Utd
Heat transfer coefficient from tedlar to heat exchanger (Wm-2K-1)
57
Vin
Input velocity of water (ms-1)
58
Greek Symbols
59
αg
Absorptivity of glass
60
αsc
Absorptivity of PV
61
αtd
Absorptivity of tedlar
62
τg
Glass transmitivity
63
εg
Glass emissivity
64
ρ
Density (kgm-3)
65
μsc
PV temperature coefficient (%/°C)
66
ηsc
Reference efficiency
67
ηe
Electrical efficiency of PV 2
ACCEPTED MANUSCRIPT 68
ηt
Thermal efficiency of PVT
69
ηo
Overall efficiency of PVT
70
1. Introduction
71
Solar energy is one of the most promising and abundant of renewable energy sources . It is of
72
great importance to develop reliable, cost-effective, and environmentally-friendly sources of
73
solar energy. Energy conversion through thermal receivers, photovoltaic (PV), and
74
photovoltaic-thermal (PVT) systems are most commonly used for solar systems. In a PVT
75
system, generally air or water is used to collect heat through a heat exchanger, which is
76
attached to the bottom tedlar surface of the PVT module [1-5]. Solar thermal collectors along
77
with PV modules combine to form the PVT system [6-10]. Numerical and experimental
78
investigations [11-13] have been conducted with various designs of PVT. In order to enhance
79
overall efficiency of PVT systems, [14-18] more focus should be given to operational factors
80
and the system design of PVT. The performance of different types of PV cell technologies
81
can vary due to different operating and weather conditions [19-22]. The PV operating
82
temperature has a significant effect on PV performance [23-26]. This performance is
83
enhanced due to reducing cell temperature. The solar spectrum was not considered to vary in
84
a relevant way in these studies.
85
The temperature of a solar cell enhances the unused energy in thegap area between cells
86
when a PV module is covered by glass at the top and tedlar at the bottom. The amount of heat
87
transfer from tedlar to ambient becomes lower for opaque PV modules than semitransparent
88
PV modules. Temperature plays an important role in PV performance and increasing
89
temperature decreases PV power by upto 7% [27-30]. The electrical power of a PV module
90
decreases with increasing temperature, while voltage reduces greatly. The impact of solar
91
radiation on the PV temperature needs to be investigated in greater detail. This investigation
92
has been performed based on the PV design with a converging lens and solar concentrator, as
93
shown in research by Avireni [31], Kerzmann and Schaefer [32], El –Sayed et al. [33] and
94
Xie et al. [34].
95
Experimental investigation of the PVT system [35] was conducted in Greece by applying a
96
dual heat extraction operation (either with water or with air circulation) for system
97
performance improvement. Numerical modelling and simulation of a hybrid PVT solar
98
energy system
99
phenomena at different locations such as Cyprus and the UK. A highly transparent low e-
[36-37] have been performed for both laminar and transient fluid flow
3
ACCEPTED MANUSCRIPT 100
coating based on silver, specifically optimized for the application in PVT collectors, was
101
developed by Laemmle et al. [38] based on experiment as well as simulation. A computer
102
simulation was performed to analyze the PVT system, facade-integrated photovoltaic/thermal
103
(BiPV/T) technology, with performance based on a dynamic model, with steady as well as
104
unsteady cases [39-40]. Based on experimental data and validated numerical models, a study
105
of the appropriateness of the glass cover on a thermosyphon-based water-heating PVT system
106
was carried out by Chow et al. [41].
107
It is found from the literature that PVT technology is a very promising field of solar energy
108
based power generation and energy conversion. Limited research has been conducted to find
109
better use of electric power generation and thermal energy from PVT modules, as well as
110
cooling systems[35-41]. However, there is a lot of scope in optimizing the outcomes of PVT
111
systems using high irradiation. In this research, a new model is developed to introduce a lens
112
(concentrator) panel on the top of an existing solar panel in order to get high irradiation. In
113
addition, a new design is implemented inside the heat exchanger using baffles to extract more
114
heat from the PVT system. The aims of the present research are to find out the effect of high
115
irradiation on PV power generation, to optimize the flow rate of cooling fluid, thermal
116
energy conversion and analysis of the overall performance of the PVT system operating
117
under high irradiation due to a lens concentrator.
118
2. Methodology
119
2.1 Experimental investigation
120 121 122 123 124 125 126 127 128 129 130 131 Figure 1: View of experimental set up where (1) Monitor, (2) Data taker, (3) MPPT, (4) Radiator, (5) Water tank, (6) Pump, (7)4 Flow meter, (8) Pyranometer, (9) Water inlet, (10) Water outlet and (11) PV module
ACCEPTED MANUSCRIPT 132
There exists an experimental setup of a SY-90M monocrystalline module in the Solar Park at
133
Level 3, Wisma R&D, UMPEDAC, University of Malaya, Malaysia. The specifications of
134
this PV module, outdoor experimental setup, instruments, test conditions, data acquisition
135
system are mentioned in Rahman et al. [2] in details. The experimental setup has been shown
136
in Figure 1, with data taken in February, 2017. The solar cell temperature was measured for
137
irradiation levels of 300 to 1000 W/m2, with a fixed cooling flow rate of 180 L/h and an
138
average inlet fluid temperature of 30°C. The obtained solar cell temperature for different
139
irradiation levels is shown in Figure 2.
140 141 142 143 144 145 146 147 148 149 150 151 152
Figure 2: Cell temperature against irradiation
153
2.2 Model development
154
The idea of using concentrator lenses is to generate electrical power more effectively using
155
the same size of solar cells [10]. In the present research, the concept of concentrating
156
irradiation has been extrapolated from the solar cell level to the PV module level. Figure 3(a-
157
c) is a schematic diagram of the PVT concentrator system in: a magnified view, full-view and
158
cross-sectional view, respectively. A lens panel (Fresnel lenses) has been implemented to
159
concentrate sunlight to every cell of the PV module. At the same time, the system usually has
160
a second optical element, namely glass, which is attached to the cell. Its mission, amongst
161
others, is to increase the numerical aperture of the set whilst at the same time ensuring a
162
uniform distribution of the energy from the cell, preventing the generation of hot spots on it.
163
Thermal paste has been used as a heat exchanger, attached to the back side of the PV module.
164
The traditional heat collection systems attached under PV cells are circular/square shaped 5
ACCEPTED MANUSCRIPT 165
parallel pipes [2] or empty box-shaped heat exchangers [42]. In traditional cases, cooling
166
fluid cannot take a high level of heat from tedlar due to it having less contact with the surface
167
of tedlar. In order to get more contact surface with cooling fluid and maximum thermal
168
energy, a new design has been developed by using sixteen baffles inside the heat exchanger.
169
In this case, water is considered as a cooling fluid. The baffle design and fluid flow path are
170
shown clearly in Figure 3(c). Heat exchangers, input-output pipes, and baffles are made of
171
aluminium metal. The four sides of the PVT module are bound by an aluminium frame and
172
the cooling system has two domains: an inlet-outlet ports-baffles of solid domain and a fluid
173
domain. In order to ensure maximum irradiation incident on the module surface, sun tracking
174
is necessary. However, in numerical simulation with COMSOL Multiphysics, the normal
175
incidence of irradiation is already given by default software settings. A large PV module of
176
72 cell-polycrystalline-silicon (with each cell area being 0.024 m2) is considered in this
177
investigation. The mean solar radiation is 1000 W/m2 in Malaysia. To get solar irradiation up
178
to 5000 W/m2, the area of each lens is 5*0.024 m2, bringing the total area of 72 lenses to
179
72*0.122 m2. The solar cell total area (1.75 m2) is considered the computational domain for
180
numerical simulation. The physical properties of different layers of the PVT module are
181
displayed in Table 1.
182
Table 1: The dimensions and properties of the PVT layers [2, 10, 42] PVT components Glass EVA Polycrystalline cell Tedlar Thermal paste Heat exchanger Inlet-outlet pipes Baffles Fluid
ρ (kg/m3) 2450 950 2329 1200 2600 237 237 237 998
Dimension (m) 2×1×0.003 2×1×0.0008 1.75×1×0.0001 2×1×0.00005 1.75×1×0.0003 1.75×1×0.012 0.05×0.01×0.01 0.978×0.001×0.01 1.75×1×0.01 (fluid region)
cp (J/kgK) 500 2090 700 1250 700 900 900 900 4200
k (W/mK) 2 0.311 148 0.15 1.9 2700 2700 2700 0.68
183
2.3 Numerical investigation
184
Considering all the outdoor experimental data (Figure 2) of a PVT system, a new correlation
185
is developed between Tsc and G:
186
where the coefficient of correlation is r2 = 97.32%. The cell temperature is calculated by
187
using the equation [2, 10]: Tsc
Tsc = 2.9692*(G)0.3948 Psc G g sc sc (U gaTamb U tTtd )
U 6
ga
Ut
(1)
(2)
ACCEPTED MANUSCRIPT Sun
188 189 Converging lens
190 191 192
Irradiation (a)
193 Solar cell
194 195 196 197 198 199 200 201 202
(b)
203 204 205 206 207 208 209 210 211 212
Glass
213Aluminium 214 frame 215
EVA Solar Cell EVA
216
Thermal paste
(c)
Tedlar
217
Outflow
218 Inflow 219 220 221
Heat exchanger
Baffle container Figure 3: Schematic diagram of (a) Fresnel lens with cell, (b) lens panel with solar panel and (c) PVT system (cross sectional view upto 7 PV layers and top view in fluid flow layer)
ACCEPTED MANUSCRIPT 222
The amount of received energy by the PV cell is found from the following equation [10] Er g sc Psc GA
223
(3)
224
A relation among concentration ratio, lens area and solar cell area [10, 33] is: Cr
225
Table 2 represents the PVT properties.
226
Al Asc
(4)
Table 2: Properties of PVT system [2, 10, 42] PVT properties Glass emissivity Transmissivity of glass Absorptivity of PV Absorptivity of tedlar PV reference efficiency PV temperature coefficient Ambient temperature Input temperature Reference temperature Packing factor Each cell area PV area Number of cells Heat transfer coefficient inside PV layers Heat transfer coefficient from tedlar to heat exchanger Heat transfer coefficient from heat exchanger to ambient Heat transfer coefficient from heat exchanger to water
Value 0.04 0.96 0.9 0.5 0.13 -0.0045 32 30 25 20% 0.156*0.156 8.76 6*12 150 77 5.84 66
227
Networks with different hidden layers have been used for 3D numerical modeling and
228
performances have been evaluated. Wind velocity is not considered and all other assumptions
229
for this numerical simulation are mentioned in [10]. The heat transfer equation for PV layers
230
[43] and laminar flow equations for the fluid domain [9-10] of a PVT system are given
231
below:
232
For the glass
233
k cp
234
For the cell
235
k cp
236
For the tedlar
2Tg 2Tg 2Tg 2 2 2 y z g x
4 4 g G U ga Tg Tamb g Tg Ts U t Tg Tsc
2Tsc 2Tsc 2Tsc 2 sc g G Ee U t Tsc Ttd U t Tsc Tg y 2 z 2 sc x
8
(5)
(6)
ACCEPTED MANUSCRIPT
237
k cp
2Ttd 2Ttd 2Ttd 2 2 x y z 2 td
238
For the heat exchanger
239
k cp
240
For the fluid domain
241
u v w 0 x y z
242
u u u p f u j v j w j f y z x x
243
f c pf u
244
where j = 1, 2, 3 in the case of u, v, w components of the velocity vector, σ
245
1.5 = 5.670367×10−8 Wm−2K−4 (Stefan-Boltzmann value), Ts 0.0552Tamb is the sky temperature.
246
Also ρf, kf, cpf and f are the density, thermal conductivity, specific heat at constant pressure
247
and dynamic viscosity of fluid, respectively. The boundary conditions of the numerical model
248
are as follows [10]:
2The 2The 2The 2 2 x y z 2 he
U t Tsc Ttd U td Ttd The
(7)
U td Ttd The U he The T f U hea The Tamb
(8)
T f T f T f v w kf y z x
(9)
2u j 2u j 2u j 2 2 2 y z x
2T f 2T f 2T f 2 2 2 y z x
249
1. for side surfaces of PVT: n . k T 0
250
2. for solid boundaries of fluid domain: u = v = w = 0
251
T T 3. for fluid-solid interface: k f khe n f n he
252
4. for inlet: T Tin , u = 0, v = Vin, w = 0
253
5. for outlet: p = 0
(10)
(11)
254
where n is the distance along x or y or z directions, acting normal to the surface.
255
The output electrical power and thermal energy of the PVT module can be calculated using
256
the following relations [10]:
257
E p sc psc g sc GA 1 sc Tsc Tr
(12)
258
and Et mc p Tout Tin
(13)
9
ACCEPTED MANUSCRIPT 259
The instantaneous electrical, thermal and overall efficiencies of the PVT system can be
260
found using the following relations, respectively [10]:
261 262 263
e
Ep Er
,
t
E Et Et and o p Er Er
(14)
The instantaneous thermal efficiency can be written another way:
FR APsc G g sc U t Tin Tamb APsc G mc p Tout Ti n
FR g sc FR U t
Tin Tamb G
264
where FR
265
value of the heat removal factor remains between 0.7 and 0.9 for a water collector system.
266
Thermal efficiency linearly depends on (Tin - Tamb) if all other terms are constant for a specific
267
PVT. Actually, the overall heat transfer coefficient, Ut, is not constant. It is a function of
268
inlet fluid and ambient air temperatures, thus it is chosen as FR U t b c Tin Tamb .
269
Then, the maximum useful energy can be rewritten as:
270 271 272 273 274
APsc G g sc U t Tin Tamb
is the PVT heat removal factor. Generally, the
Qu sfl FR APsc G g sc bAPsc Tin Tamb cAPsc Tin Tamb
2
Thus, the derived quadratic form of thermal efficiency is:
FR APsc G g sc bAPsc Tin Tamb cAPsc Tin Tamb Q usfl AI APsc G
T T T T FR g sc b in amb c in amb G
G
2
2
a bT * cGT *2
(15)
275
where a = FR τg αsc and T* = (Tin - Tamb)/G.
276
2.4. Numerical Modeling
277
The finite element method (FEM) with Galerkin's weighted residual technique based on
278
software COMSOL Multiphysics is used for solving the 3D numerical model of PVT. It is a
279
finite element analysis solver and simulation software for various physics and engineering
280
applications, especially coupled phenomena, or multiphysics. This software is also used for
281
creating physics-based applications. It allows entering coupled systems of partial differential
282
equations (PDEs). The PDEs can be entered directly or using the so-called weak form. The
283
finite element technique [44] is applied in order to solve the system of Equations (5)-(11) for
284
the present simulation. The numerical technique is described in detail in Nasrin and Alim 10
ACCEPTED MANUSCRIPT 285
[45]. Tecplot and Microsoft-Excel are software used to plot the graphs by exporting
286
simulated data from COMSOL Multiphysics.
287
2.4.1 Meshing and grid test
288
The finite element meshing of the computational domain of a PVT module is displayed in
289
Figure 4. In the present numerical model, the subdomain and boundary elements have been
290
chosen as free tetrahedral and free triangular forms, respectively. At irradiation level of 1000
291
W/m2 and a flow rate of 180 L/h have been chosen, and a grid test has been conducted for the
292
PVT model. Different types of non-uniform grid systems are checked with elements:
293
2,68,882; 4,28,585; 8,13,304; 14,34,582; and 32,11,718. Supervising parameters are chosen
294
as cell temperature and outlet fluid temperature. It was noticed that there was no considerable
295
change in the value of cell and outlet temperatures between normal and fine meshing but time
296
intolerable. Thus, the PVT model with 14,34,582 domain elements is considered for
297
numerical analysis. The grid test result is depicted in Table 3.
298
Table 3: Grid test at irradiation 1000 W/m2 and volume flow rate 180 L/h Meshing type
Extra Coarse
Coarser
Coarse
Normal
Fine
No. of elements
2,68,882
4,28,585
8,13,304
14,34,582
32,11,718
Tsc (°C)
41.8535
41.3248
42.8751
43.0112
43.0118
Tout (°C)
34.0214
34.5321
35.0634
35.4521
35.4526
Time (s)
719
895
1081
1225
1831
299 300 301 302 303 304 305
Figure 4: Finite element meshing of PVT model
306
2.4.2. Simulated steady-state condition
307
The solar radiance, ambient temperature, wind velocity, etc. can be varied any time and
308
depend on the operating weather conditions. Thus, PVT operation is inherently dynamic. The 11
ACCEPTED MANUSCRIPT 309
steady-state condition is required in numerical simulation for producing accurate results and
310
this is explained in more detail in Nasrin et al. [10]. For this purpose, the numerical
311
simulation has been performed at fixed a inlet temperature of 30°C, with an irradiation level
312
of 1000 W/m2 and a flow rate of 180 L/h. Figure 5 expresses the values of cell and output
313
temperatures against operating time. The steady-state condition was reached after 1200 s.
314 315 316 317 318 319 320 321 322 323 324
Figure 5: Steady-state condition in 3D model simulation
325
2.4.3. Code validation
326
The validity of present 3D simulation has been conducted with experimental as well as
327
numerical results which are as follows:
328
2.4.3.1 Validity test with experimental findings
329
The values of the solar cell temperature at an irradiation level of 1000 W/m2 with various
330
flow rates (30, 60, 90 and 180 L/h) was obtained from the present numerical code, validated
331
with that of Rahman et al. [2]. The authors [2] used an SY-90M PV (1200 * 545 * 35 mm)
332
module of six layers, with 4*9 monocrystalline cells, a rectangular heat exchanger (950 * 420
333
mm) composed of seven copper tubes (each having a 22 mm diameter) and with an inlet
334
temperature of 35°C. Table 4 expresses this validation and represents a very good agreement
335
with experimental results [2].
336
2.4.3.2 Validity test with numerical findings
337
The values of surface temperature of the PVT system with an inlet fluid temperature of 34°C,
338
solar radiation of 1000 W/m2, and an inlet velocity of 0.0007 m/s has been obtained using the
339
present numerical code, validated with that of Nahar et al. [42]. Figure 6 displays the
12
ACCEPTED MANUSCRIPT 340
validation and a good match of results is found. This code validation is described in detail in
341
Nasrin et al. [10].
342
Table 4: Code validation of cell temperature against flow rate Mass flow rate (L/h) 30 60
Cell temperature (°C) Present research Rahman et al. [2] 51.11 52.88 48.04 50.23
Percentage of error 3.3% 4.3%
90
47.70
49.65
3.9%
180
45.76
47.76
4.2%
343 344
126
345 346 347 348 349 350 351
Present code Nahar et al. [42] Figure 6: Model validation of surface temperature of PVT system
352 353
3. Results and Discussion
354
The present research is conducted to analyze the effects of high irradiation and volume flow
355
rate on PVT performance. The variety of solar radiation and volume flow rate of inlet fluid
356
has been chosen as 1000 to 5000 W/m2 and from 30 to 210 L/h, respectively. The outcomes
357
of the different cases are presented in the following sections.
358
3.1 Effect of irradiation
359
Figure 7 depicts the solar cell surface temperature along the module length at the middle
360
section (width 500 mm) with varying levels of solar irradiance (1000, 2000, 3000, 4000 and
361
5000 W/m2). In this figure, the cooling system’s flow rate was 180 L/h. It is observed from
362
this figure that the solar cell temperature of the PVT module increases along the PV length at
363
the initial solar irradiation level of 1000 W/m2. The variation of the cell temperature of the
364
PV module becomes higher for increasing values of irradiation. Equation (2) shows that solar
365
cell temperature is linearly proportional to solar radiation. This is justified by the heat
366
transfer procedure of the PVT system. 13
ACCEPTED MANUSCRIPT 367 368 369 370 371 372 373 374 375
Figure 7: Cell temperature along module length for different irradiation
376 377
The streamlines have been plotted for the PVT system at different values of irradiation from
378
1000 to 5000 W/m2 with a flow rate of 180 L/h, as shown in Figure 8. The average
379
temperature of the water at the exit port becomes higher with increasing values of irradiation.
380
This is due to the fact that increasing solar radiation increases the surface temperature of the
381
PV module, resulting in more heat being produced by the solar cell. Consequently, more heat
382
is absorbed by the tedlar surface and the heat transfer system from the solar cell to the
383
cooling fluid through the heat exchanger is enhanced. As a result, the outlet fluid’s mean
384
temperature becomes higher.
385
3.2 Effect of flow rate
386
Figure 9 depicts the effect of flow rate from 30 to 210 L/h on the surface temperature plot for
387
the PVT system at an irradiation of 1000 W/m2. The minimum and maximum temperatures
388
for each surface temperature plot are represented by blue and red colors, respectively. This
389
figure shows that the rising volume flow rate accelerates thermal current activities through
390
the heat exchanger surface to the fluid flow domain. With increasing flow rate, the color of
391
the surface near the outlet port of the PVT module becomes lighter, whereas initially this
392
color becomes deeper. With the variation of flow rate from 30 to 210 L/h, the temperature
393
distribution becomes distorted, resulting in an increase in the overall heat transfer. It is found
394
that increasing the value of the inlet fluid flow rate results in the maximum temperature of the
395
PVT material dropping gradually. At the lowest flow rate (30 L/h), increasing temperature is
396
the maximum and at the highest flow rate (210 L/h), it is minimized. 14
ACCEPTED MANUSCRIPT 397 398 399
5000 W/m2
400 401 402 403 404 405 406 407
4000 W/m2
408 409 410 411 412 413 414
3000 W/m2
415 416 417 418 419 420 421
2000 W/m2
422 423 424 425 426 427 428 429
1000 W/m2
430 15various irradiation at flow rate 180 L/h Figure 8: Streamlines of water for
ACCEPTED MANUSCRIPT 431 432 433 434
210 L/h
435 436 437 438 439 440 441
180 L/h
442 443 444 445 446 447 448 449
90 L/h
450 451 452 453 454 455 456
60 L/h
457 458 459 460 461 462 463
30 L/h
464 16 Figure 9: PVT surface temperature for various flow rate at irradiation 1000 W/m2
ACCEPTED MANUSCRIPT 465 466 467 468
210 L/h
469 470 471 472 473 474 475
180 L/h
476 477 478 479 480 481 482
90 L/h
483 484 485 486 487 488 489
60 L/h
490 491 492 493 494 495 496 497
30 L/h
498 17
Figure 10: Streamlines of water for various flow rate at irradiation 1000 W/m2
ACCEPTED MANUSCRIPT 499
This illustrates the dominating behavior of fluid flow properties. Thus, conductive as well as
500
convective heat transfer processes run from the top glass surface to the outlet exit through the
501
heat exchanger of the PVT module.
502
The streamlines of water for different values of flow rate (30, 60, 90, 180 and 210 L/h) are
503
displayed in Figure 10. For this simulation, solar radiation is chosen as 1000 W/m2. Figure 10
504
shows the fluid flow distribution throughout the heat exchanger from inlet to outlet. When the
505
volume flow rate is low, the temperature of streamlines at the exit port becomes high. The
506
outlet fluid temperature reduces with increasing values of the flow rate. This result is
507
significant because rapidly flowing fluid is not capable of taking in more heat from the heat
508
exchanger. So, the volume flow rate of the inlet fluid is very important in the cooling system
509
of the PVT module. The pumping power for the volume flow rate at 210 L/h is higher than
510
other flow rates.
511
3.3 Solar cell temperature
512
Figure 11(a)-(b) depicts the solar cell temperature against various solar radiation levels and
513
fluid flow rates, respectively. Increasing the rate of cell temperature by 0.81% increases
514
irradiation from 1000 to 5000 W/m2 in the PVT system. Teo et al. [1], Rahman et al. [2],
515
Nasrin et al. [10] and Chandrasekar et al. [46] found a 1.4, 2.71, 1.85 and 1.4°C increase in
516
cell temperature for every 100 W/m2 increase in the irradiation level with the cooling system.
517
The operating irradiation level of Teo et al. [1], Rahman et al. [2], Nasrin et al. [10] and
518
Chandrasekar et al. [46] were 550 to 1050 W/m2, 240 to 979 W/m2, 1000 to 3000 W/m2 and
519
600 to 1300 W/m2, respectively. The present result shows that the cell temperature increases
520
about 0.9°C for each increase of irradiation 100 W/m2. The aim of this research is to maintain
521
the cell temperature under a certain range using the proper cooling system so that the PV
522
material may not be degraded. This result is better than that of Teo et al. [1], Rahman et al.
523
[2], Nasrin et al. [10] and Chandrasekar et al. [46]. This implies that the proper cooling
524
system is maintained in this PVT system.
525
Figure 11(b) shows that the average cell temperature, Tsc decreases with increasing inlet fluid
526
volume flow rate from 30 to 210 L/h at the irradiation level of 1000 W/m2, and with an inlet
527
fluid temperature of 30°C. As the inlet fluid flow rate increases, more heat is removed from
528
the module by convection which reduces the average cell temperature. It is found from this
529
figure that the cell mean temperature decreases rapidly with increasing flow rate. Initially, the
530
flow rate is 30 L/h, with a solar cell temperature of 67°C. The cell temperature decreases 18
ACCEPTED MANUSCRIPT 531
gradually as the flow rate increase up to 180 L/h. After that, at a flow rate of 210 L/h, it
532
reduces a little bit to 42.6°C, but the pumping power is higher. Thus, for this PVT system, the
533
optimum volume flow rate of cooling fluid is 180 L/h. The solar cell temperature decreases
534
by 1.1°C for each 10 L/h increment of volume flow rate of inlet fluid.
535 536 537 538 539 540 541 542 543 544 545 546 547
(b) (a) Figure 11: Solar cell temperature with the variation of (a) irradiation at flow rate 180 L/h and (b) flow rate at irradiation 1000 W/m2
548
Equation (1) shows the correlation of cell temperature and irradiation from the experimental
549
result for low irradiation (300 - 1000 W/m2). Another correlation was developed from the
550
result of the numerical simulation for high irradiation (1000 - 5000 W/m2) at the fixed
551
cooling fluid flow rate of 180 L/h. This is written as:
552
Tsc = 3.4913*(G)0.3641
(16)
553
where the coefficient of correlation is r2 = 99.02%. Equation (16) has a similar pattern as
554
Equation (1). This demonstrates numerical validity. The percentage of error between these
555
two correlations for the PV cell temperature at an irradiation of 1000 W/m2 is less than 5%.
556
3.4 Electrical power
557
The output power for the variation of irradiation (1000 - 5000 W/m2) and flow rate (30 - 210
558
L/h) is expressed in Figure 12(a)-(b). It is seen from Figure 12(a) that the initial irradiation
559
level is 1000 W/m2, the output power is 197 W, the irradiation level is 5000 W/m2 and the
560
electrical power is 983 W. For every 100 W/m2 increment in irradiation, there is an increase
561
of 19.65 W in electrical power of the PVT system at a flow rate of 180 L/h. Rahman et al.
562
[2], Nasrin et al. [10] and Nahar et al. [42] showed that the output power increased about 19
ACCEPTED MANUSCRIPT 563
3.88, 6.4 and 10 W, respectively, for every 100 W/m2 increase in irradiation level. So, the
564
present numerical result is consistent with these authors [42] but differs a little bit from
565
Rahman et al. [2] and Nasrin et al. [10]. This occurs because of the size of the PV module,
566
the packing factor, the reference temperature, operating conditions, poor installation of the
567
cooling system and the properties of materials used.
568
At a fixed irradiation level of 1000 W/m2, under different volume flow rates of fluid, the
569
initial flow rate is 30 L/h and the electrical power is 195 W. During the peak flow rate of 180
570
L/h, the output power becomes 197 W. But there is no significant increase in output power
571
(197.09 W) for further increases of water flow rate upto 210 L/h. The output power increased
572
by 2.09 W with an approximately 180 L/h increase in the volume flow rate of the fluid.
573
Under cooling conditions, the output power increased by 0.12 W as for each 10 L/h
574
increment of fluid flow rate.
575 576 577 578 579 580 581 582 583 584 585 586 587
(a)
(b)
Figure 12: Electrical power versus (a) irradiation at flow rate 180 L/h and (b) flow rate at irradiation 1000 W/m2
588
3.5 Electrical efficiency
589
The electrical efficiency of the PVT system as a function of the solar irradiance and volume
590
flow rate varies from 1000 to 5000 W/m2 and 30 to 180 L/h, respectively, as shown in Figure
591
13(a)-(b). The PV efficiency decreases with rising irradiation level at the flow rate of 180
592
L/h. The electrical efficiency devalues from 13 to 10.6% due to increasing solar radiation. So,
593
electrical efficiency decreases approximately 0.06% for each 100 W/m2 increment of
594
irradiation. Rahman et al. [2], Nasrin et al. [10] and Nahar et al. [42] showed that PV
20
ACCEPTED MANUSCRIPT 595
efficiency decreased about 0.87%, 0.09% and 0.16% for every 100 W/m2 increase of
596
irradiance.
597
Due to increasing inlet volume flow rate (30 - 210 L/h) of water, the cell average temperature
598
(Figure 11(b)) is reduced. Consequently, the PVT module’s current drops marginally with a
599
noticeable increase in PVT voltage which, in turn, increases the output power and electrical
600
efficiency. The maximum electrical efficiency was found to be around 13.05%. There wa no
601
noticeable increment observed for increasing the flow rate from 180 to 210 L/h. So, for the
602
cooling system of the PVT module, an inlet fluid flow rate not more than 180 L/h is
603
advantageous. Electrical efficiency increases by approximately 0.03% for each increase of
604
fluid flow rate by 10 L/h.
605 606 607 608 609 610 611 612 613 614 615 616
(b) (a) Figure 13: Electrical efficiency versus (a) irradiation at flow rate 180 L/h and (b) flow rate at irradiation 1000 W/m2
617 618
3.6 Outlet temperature of cooling fluid
619
Figure 14(a)-(b) shows the effect of G (1000 - 5000 W/m2) and flow rate (30 - 210 L/h) on
620
mean temperature of the outlet cooling fluid. Rising solar radiance enhances outlet
621
temperature. At the fixed volume flow rate of 180 L/h and irradiation of 1000 W/m2, the
622
outlet fluid temperature was found to be approximately 35.45oC. This outlet temperature
623
became 55.2oC at the highest irradiation level of 5000 W/m2. So, the temperature of the outlet
624
fluid was enhanced by 19.75oC for each 4000 W/m2 increment of irradiation. The outlet
625
temperature increases about 0.5oC for each 100 W/m2 increase of solar radiance.
626
The water outlet temperature also decreases with increasing inlet flow rate at G = 1000 W/m2,
627
as shown in Figure 14(b). This declining trend can also be attributed to the increasing 21
ACCEPTED MANUSCRIPT 628
convection heat transfer rate with increasing velocity. With the increase in flow velocity, the
629
rate of heat removal is also increased and less time is available for thermal accumulation,
630
thereby decreasing the water outlet temperature. Nahar et al. [42] represented the
631
experimental result where the graph of outlet fluid temperature against inlet velocity had a
632
similar trend, but different values were presented due to different inlet velocity and
633
conditions. The similarity in trend and proximity in magnitude between the present numerical
634
value and that of Nahar et al. [42] validates the present numerical model with quiet
635
confidence. This decreasing pattern changes significantly up to a 180 L/h flow rate of cooling
636
water. It slightly decreases as the flow rate increases and exceeds 180 L/h. Hence, 180 L/h is
637
the most suitable flow rate to enhance the efficiency of the PV module.
638 639 640 641 642 643 644 645 646 647 648 649 650
(a) (b) Figure 14: Outlet temperature against (a) irradiation at flow rate 180 L/h and (b) flow rate at irradiation 1000 W/m2
651
3.7 Thermal energy
652
The thermal energy with respect to various irradiation and volume flow rates is displayed in
653
Figure 15(a)-(b). At the flow rate 180 L/h and G = 1000 W/m2, the thermal energy was found
654
to be 1165 W. This energy became 5387 W for G = 5000 W/m2. The reason for this fact is
655
that there is conductive heat transfer inside the PVT module’s surfaces, as well as convective
656
heat transfer inside the fluid. Therefore, a high temperature gradient takes place between the
657
output and input fluids due to high irradiation. The thermal energy increases by 105.55 W for
658
every increase of irradiance by 100 W/m2. This huge amount of thermal energy can be used
659
for secondary purposes such as cooking, cleaning, bathing, drying, pain removal, space
22
ACCEPTED MANUSCRIPT 660
heating, building, ventilation etc. in households as well as in industries and agricultural
661
sectors.
662
At the irradiation level of 1000 W/m2, Figure 15(b) shows that the total amount of thermal
663
energy (1055 W) has to be collected from the system when the flow rate is 30 L/h. At the
664
flow rate of 180 L/h, thermal energy is found to be 1165 W and at the highest flow rate of
665
210 L/h, the amount of thermal energy becomes 1170 W. Only 5 W of extra thermal energy
666
is obtained for further increasing the flow rate up to 210 L/h. Actually, increasing the flow
667
rate of the cooling fluid from 30 to 60 L/h significantly affects the decreasing rate of outlet
668
fluid temperature as shown in Figure 14(b). After that, increasing the flow rate upto 180 L/h
669
decreases the outlet fluid temperature gradually. Nonetheless, at a flow rate of 210 L/h, the
670
output temperature decreases very slowly. Consequently, the extracted thermal energy from
671
the system enhances slowly when the cooling fluid’s flow rate exceeds 180 L/h. Thus, the
672
optimum flow rate of inlet fluid is 180 L/h for proper cooling of the PVT. For each 10 L/h
673
increment of volume flow rate, the extracted heat energy is 6.4 W.
674 675 676 677 678 679 680 681 682 683 684 685 686
(a) (b) Figure 15: Thermal energy against (a) irradiation at flow rate 180 L/h and (b) flow rate at irradiation 1000 W/m2
687
3.8 Thermal efficiency
688
The rate of thermal efficiency decreases for various values of irradiation from 1000 to 3000
689
W/m2 in (Figure 16(a)) because the total amount of receiving energy from the PVT system is
690
increased for escalating irradiation. The thermal efficiency of the PVT system devalues from
691
77% to 71% for increasing solar radiance. For every increase of 100 W/m2, irradiation
23
ACCEPTED MANUSCRIPT 692
thermal efficiency decreases to about 0.15%. Nasrin et al. [10] found a 0.3% decrement of
693
thermal efficiency for each 100 W/m2 increment of irradiation.
694
The thermal efficiency increase with increasing volume flow rate of the working fluid at a
695
fixed irradiation of 1000 W/m2 is shown in Figure 16(b). An increase in flow rate from 30 to
696
180 L/h enhances the convective heat transfer coefficient of the cooling fluid. So, under a
697
given temperature difference, more heat is transferred at higher velocities, thereby increasing
698
the thermal efficiency. The thermal efficiency increases from 70% to 77% for this numerical
699
simulation. The efficiency rate is high due to the variation of the fluid flow rate from 30 to
700
180 L/h. After that, at a flow rate of 210 L/h, a very small increment in thermal efficiency is
701
observed. The fact behind this trend is that increasing the volume flow rate (up to 180 L/h) of
702
cooling fluid decreases solar cell temperature gradually. The solar cell temperature slightly
703
decreases as the flow rate increases and exceeds 180 L/h. Hence, 180 L/h is the most suitable
704
flow rate to enhance the efficiency of the PVT system. About 0.43% thermal efficiency
705
increases due to each 10 L/h increase of volume flow rate of inlet water.
706 707 708 709 710 711 712 713 714 715
(b) (a) Figure 16: Thermal energy versus (a) irradiation at flow rate 180 L/h and (b) flow rate at irradiation 1000 W/m2
716
3.9 Overall efficiency
717
The PVT overall efficiency versus irradiation (1000 - 5000 W/m2) and flow rate (30 - 210
718
L/h) is expressed in Figure 17(a)-(b). It decreases from 90% to 81.6% due to rising irradiance
719
at the flow rate of 180 L/h. The overall efficiency of the PVT system devalues by 0.21% for
720
each increase of 100 W/m2 irradiation. Nasrin et al. [10] showed that the overall efficiency
721
decreased by 0.39% due to each increase of 100 W/m2 irradiation. 24
ACCEPTED MANUSCRIPT 722
As both the electrical and thermal efficiencies increase with increasing inlet fluid flow rate,
723
so the overall efficiency also increases. Accordingly, at lower flow rates, overall efficiency
724
was obtained as 82.5%. It increased significantly up to 90% with increasing flow rates up to
725
180 L/h. After that, the value of overall efficiency for the volume flow rate of 210 L/h is
726
approximately 90.85%. So, a 180 L/h flow rate is perfect for this PVT module. Overall,
727
efficiency increases about 0.46% for each 10 L/h increase of volume flow rate.
728 729 730 731 732 733 734 735 736 737 738 739 740
(a) (b) Figure 17: Overall efficiency against (a) irradiation at flow rate 180 L/h and (b) flow rate at irradiation 1000 W/m2
741
3.10 Comparison
742
3.10.1 PV temperature
743
The increasing rate of cell temperature for every 100 W/m2 increase in irradiation resulting
744
from various research such as Teo et al. [1], Rahman et al. [2], Nasrin et al. [10],
745
Chandrasekar et al. [46], Bahaidarah et al. [47] and the current work is shown in Table 5. A
746
good agreement was noticed between the present result and Teo et al. [1], Nasrin et al. [10],
747
Chandrasekar et al. [46], Bahaidarah et al. [47]. Nevertheless, the result is a little different
748
from Rahman et al. [2] because of the variation in the range of incident solar irradiation, the
749
range of the module’s operating temperature, wind velocity, ambient temperature, and
750
physical properties of each layer of the PVT module.
751
Table 5: Comparison of cell temperature increment per 100 W/m2 irradiance Investigations
G (W/m2) Starting Ending
Tsc (°C) Tsc increment per 100 Tamb W/m2 irradiance (°C) Starting Ending 25
ACCEPTED MANUSCRIPT [1] [2]
550 312
1050 995
41 31
48 50
1.4 2.71
35
[10] [46]
1000 600
3000 1300
48 40
85 50
1.85 1.4
32 37
[47]
240
979
21
35
1.9
21
Present research
1000
5000
43
78
0.9
32
752 753
3.10.2 Electrical efficiency
754
The obtained numerical results were compared for electrical efficiency with the variation of
755
flow rate with the experimental result of Rahman et al. [2]. For this comparison, input fluid
756
temperature was 33°C, ambient temperature was 35°C, and solar radiation was 996 W/m2.
757
These parameters were kept fixed. The experiment of Rahman et al. [2] was operated to find
758
the energy and efficiency of the PVT module in Malaysia. Figure 18 demonstrates the above
759
stated comparison. The percentage of error between numerical and experimental data lies
760
between 0 to 5%. This is due to the different operational factors at the time of experimental
761
investigation, such as ambient temperature, reference temperature, wind velocity, and the
762
effect of dust and other factors. In numerical investigations, the input and ambient
763
temperatures are fixed, but in outdoor experiments such as that of Rahman et al. [2], the
764
fluctuation of ambient and inlet water temperature andalso wind speed could not be
765
controlled.
766 767 768 769 770 771 772 773 774 775 776
Figure 18: Comparison for electrical efficiency between present result and [2]
777 778
3.10.3 Output temperature and thermal efficiency 26
ACCEPTED MANUSCRIPT 779
The variations of output fluid temperature and thermal efficiency for increasing values of
780
input fluid velocity are observed from this numerical simulation and then compared with the
781
experimental result of Nahar et al. [42]. The comparison is displayed in Figure 19(a)-(b). The
782
PVT model is solved numerically with water as the working fluid, the input fluid temperature
783
at 34°C, ambient temperature at 34°C, and irradiation at 1000 W/m2. The result of the present
784
investigation of a PVT system is little different from Nahar et al. [42] because of the
785
variation of module operating temperature, wind velocity, reference temperature, sky
786
temperature, ambient temperature and other conditions. The numerical investigation was
787
conducted with constant irradiation, reference temperature, sky temperature, ambient
788
temperature, etc. It was noticed that although the experimental curve followed the similar
789
trend of numerical results, the numerical values were somewhat higher than the
790
corresponding experimental values. This discrepancy between the numerical and
791
experimental values is due to the uncontrollable outdoor ambient conditions of the
792
experimental site where an inlet temperature of water could not be maintained at a certain
793
level as in the case of numerical simulation. So, the reason behind these incongruities is due
794
to the unrestrained ambient conditions that prevail in an outdoor experiment.
795 796 797 798 799 800 801 802 803 804 805 806 807
(a) (b) Figure 19: Comparison between numerical and experimental [42] data for (a) output temperature and (b) thermal efficiency
808 809
3.10.4 Quadratic regression equation
810
A quadratic form of the regression equation is developed for PVT thermal efficiency (ηt) and
811
reduced temperature T* (Km2/W) by using Equation (15). The regression equation is: 27
ACCEPTED MANUSCRIPT 812
ηt = - 0.0873 T*2 - 7.0209 T* + 0.6035
(17)
813
where the confidence coefficient is r² = 0.9805.
814
This quadratic regression equation is calculated in the range of reduced temperature
815
Tout Tin Tin 2 - 0.01 < T* < 0.055, where T * G
Tamb .
816
The graphical form of this regression equation is depicted in Figure 20. Generally, the slope
817
becomes lower due to reduced temperatures as the electrical efficiency of PV cells decreases,
818
resulting in a relatively higher thermal efficiency of PVT module.
819 820 821 823 824
ηt
822
825 826 827 828 829
T* (Km2/W) Figure 20: Regression graph for thermal efficiency against reduced temperature
830 831
Bosanac et al. [48] conducted an experimental and numerical analysis of a PVT solar
832
collector in Denmark. The authors [48] used crystalline Si cells, with a solar irradiation level
833
of 800 W/m2, a variation of inlet fluid temperature from 10 to 50°C, a wind velocity of 5 m/s,
834
an ambient temperature of 20°C, a PV area of 2.2 m2, emittance of PV cells at 0.95, a PV
835
reference temperature of 25°C, with the power temperature coefficient of PV cells being
836
0.40%/K. The regression equation of [48] was measured as:
837
ηt = -18.338T*2 - 6.4965T* + 0.6193 with confidence coefficient r2 = 0.988.
838
The comparison for quadratic regression equations between the present result (Equation (17))
839
and Bosanac et al. [48] is expressed in Table 6. For this comparison, the present work is
840
conducted by considering solar irradiation levels of 800 W/m2, a variation of inlet fluid
841
temperature from 10 to 50°C, an ambient temperature of 20°C. At the reduced temperature,
28
ACCEPTED MANUSCRIPT 842
the computed error between the present numerical result and that of Bosanac et al. [48] is
843
4.4%.
844
Table 6: Comparison of regression equations Efficiency at T* = 0.025 Km2/W Error
Investigations
Quadratic regression equations
[48]
ηt = - 18.338T*2 - 6.4965T* + 0.6193
45%
Present work ηt = - 0.0873T*2 - 7.0209T* + 0.6035
43%
4.4%
845 846
3.10.5 Quadratic form of thermal efficiency
847
Figure 21 shows the comparison of thermal efficiency based on quadratic form (Equation
848
(15)) against reduced temperature, T* (Km2/W) between the present simulated numerical
849
result and that of Lammle et al. [38]. In their numerical model, they used c-Si cells, low-
850
emmisivity coatings on glass surface, a solar irradiation level of 1000 W/m2, a mass flow rate
851
of 90 Kg/h, a linear heat loss coefficient of 6.37 W/m2K, PV reference electrical efficiency at
852
11.5%, an inlet fluid temperature of 25°C, a wind velocity of 3 m/s, an ambient temperature
853
of 25°C, with emissivity of PV cells at 0.89, and with the temperature coefficient of PV cells
854
being 0.54%/K. The comparison was shown for without coatings and a good agreement
855
between the present result and that of Lammle et al. [38] is observed.
856 857 858 859 860 861 862 863 864 865 866 867 868
T* (Km2/W) Figure 21: Comparison for thermal efficiency against reduced temperature
869
29
ACCEPTED MANUSCRIPT 870
Again, the comparison of the percentage of thermal efficiency against reduced temperature,
871
T* (Km2/W) between current numerical result and that of Guarracino et al. [39] is displayed
872
in Figure 22. In their numerical model, they conducted electrical-thermal modelling of sheet-
873
tube hybrid photovoltaiv/thermal collectors. The comparison is conducted using single glazed
874
collector, solar irradiation level of 1000 W/m2, linear heat loss coefficient of 4.028 W/m2K,
875
PV reference electrical efficiency of 12.6%, 14 pipes, inlet fluid temperature of 20°C, wind
876
velocity of 1 m/s, nominal electrical power of 180 W, ambient temperature of 20°C, PV area
877
of 1.43 m2 with 72 cells, with emissivity of PV cells being 0.90, transmitivity of glazing
878
being 0.95, and the temperature coefficient of PV cells being 0.40%/K. A difference occurred
879
due to the novel design of the present PVT system, though the pattern of the graph of thermal
880
efficiency against reduced temperature is similar.
881 882 883 884 885 886 887 888 889 890 891 892 893
T* (Km2/W) Figure 22: Compared results for thermal efficiency against reduced temperature
894 895
3.11 Correlation
896
From the results of this numerical simulation, a correlation was developed among solar cell
897
temperature, irradiation, ambient temperature, cooling fluid input temperature and volume
898
flow rate of cooling fluid. The correlations were conducted for solar irradiation levels from
899
1000 to 5000 W/m2, cooling fluid volume flow rates from 30 to 210 L/h, and ambient air and
900
inlet fluid temperatures 32ºC and 30ºC respectively. This can be written as:
901
Tsc = (0.2491*Tamb + 0.12736*Tin) (G)0.3691 (V)-0.245
902
where the correlation coefficient is r2 = 0.9881. 30
(18)
ACCEPTED MANUSCRIPT 903
4. Conclusion
904
The investigation shows that high solar radiation and thevolume flow rate of cooling fluid
905
have significant effects on PVT performance. Various irradiation levels and flow rates have
906
been applied for getting high performances of PVT by maintaining optimum cooling systems.
907
The electrical power generation and thermal energy conversion of the PVT system increases
908
with increasing irradiation and varying cooling fluid flow rate. The following conclusions can
909
be drawn from the present research:
910
The optimum cooling fluid flow rate is found to be 180 L/h for a PVT module exposed to high irradiation of up to 5000 W/m2.
911 912
For every 100 W/m2 increase in solar radiation, the cell temperature, outlet fluid
913
temperature, electrical power and thermal energy increase approximately 0.9, 0.5ºC and
914
19.65, 105.55 W, respectively.
915
for each 100 W/m2 increase of irradiation, respectively.
916 917
The electrical, thermal and overall efficiencies decrease about 0.06%, 0.15% and 0.21%
The cell temperature and outlet fluid temperature decrease approximately 1.1 and 1.4ºC
918
respectively; electrical power and thermal energy increase about 0.12 and 6.4 W,
919
respectively, for every 10 L/h fluid flow rate increase.
920 921
For every 10 L/h increment of
fluid flow rate, the electrical, thermal and overall
efficiencies increase about 0.03%, 0.43% and 0.46%, respectively.
922
In this present research, a novel model was developed by using the converging lens panel on
923
the solar PV panel as well as a new design of heat exchanger (aluminium metal) consisting
924
sixteen baffles is attached directly to the PV module. The proposed model is simple and
925
suitable for building integration, providing air/water depending on the season, and for the
926
thermal requirements of the building, etc. Using high irradiation, this model is able to
927
generate more electrical power and extract thermal energy. The present developed model was
928
found to perform well enough in enhancing the overall efficiency of the PVT system.
929
Acknowledgement
930
Using the economic support of UMPEDAC, HICoE grant, Ministry of Higher Education
931
(Project: UM.0000067/HME.OM, UMPEDAC - 2016) the present research has been done.
31
ACCEPTED MANUSCRIPT 932
References
933
1.
934 935
Teo HG, Lee PS, Hawlader MNA, An active cooling system for photovoltaic modules, Appl. Energy, 2012; 90(1): 309-315.
2.
Rahman, M.M., Hasanuzzaman, M., Rahim, N.A., Effects of operational conditions on
936
the energy efficiency of Photovoltaic modules operating in Malaysia, J. of Cleaner
937
Production, 2017; 143: 912-924.
938
3.
Ibrahim A., Othman M.Y., Ruslan M.H., Alghoul M.A., Yahya, M. and Zaharim, A. and
939
Sopian, K., Performance of photovoltaic thermal collector (PVT) with different
940
absorbers design, WSEAS Trans. on Environ. and Develop., 2009; 5(3): 321-330
941
4.
942 943
Chow T.T., A review on photovoltaic/thermal hybrid solar technology, Appl. Energy, 2010; 87: 365–379.
5.
Koech, R.K., Ondieki, H.O., Tonui, J.K., Rotich, S.K., A steady state thermal model for
944
Photovoltaic/Thermal (PV/T) system under various conditions, Int. J. of Sci. & Tech.
945
Res., 2012; 1: 11.
946
6.
Nasrin, R., Alim, M.A. and Chamkha, A.J., Effects of physical parameters on natural
947
convection in a solar collector filled with nanofluid, Heat Trans.-Asian Res., 2013; 42(1):
948
73-88.
949
7.
Jin G.L, Ibrahim A., Chean Y.K, Daghigh R., Ruslan H., Mat S. Othman M.Y and
950
Sopian K., Evaluation of single-pass photovoltaic-thermal air collector with rectangle
951
tunnel absorber, Amer. J. of Appl. Sci., 2010; 2: 277-282.
952
8.
953 954
Nasrin, R. and Alim, M.A., Semi-empirical relation for forced convective analysis through a solar collector, Solar Energy, 2014; 105: 455-467.
9.
Nasrin, R., Parvin, S. and Alim, M.A., Heat transfer and collector efficiency through
955
direct absorption solar collector with radiative heat flux effect, Num. Heat Trans. Part A-
956
Application, 2015; 68: 1-21.
957
10. Nasrin, R., Hasanuzzaman, M. and Rahim, N.A., Effect of high irradiation on
958
photovoltaic power and energy, Int. J. of Energy Res., 2017; DOI:10.1002/er.3907
959
(Article in press).
32
ACCEPTED MANUSCRIPT 960
11. Kalogiroua, S.A. , Karellas, S., Badescu, V., Braimakis, K., Exergy analysis on solar
961
thermal systems: A better understanding of their sustainability, Renew. Energy, 2016;
962
85: 1328-1333.
963
12. Hasanuzzaman M, Al-Amin AQ, Khanam S, Hosenuzzaman M, Photovoltaic power
964
generation and its economic and environmental future in Bangladesh, J. of Renew. and
965
Sust. Energy, 2015; 7(1): 013108.
966
13. Tiwari A, Sodha MS, Parametric study of various configurations of hybrid PV/thermal
967
air collector: Experimental validation of theoretical model, Solar Energy Mat. & Solar
968
Cells, 2007; 91:17–28.
969
14. Ahmed F, Al Amin AQ, Hasanuzzaman M, Saidur R., Alternative energy resources in
970
Bangladesh and future prospect, Renew. and Sust. Energy Rev., 2013; 25: 698-707.
971
15. Joshi AS, Tiwari A, Tiwari GN, Dincer I, Reddy BV, Performance evaluation of a hybrid
972
photovoltaic thermal (PV/T) (glass-to-glass) system, Int J. Therm Sci., 2009; 48:154–64.
973
16. Sarhaddi F., Farahat S., Ajam H., Behzadmehr A., Adeli M.M., An improved thermal
974
and electrical model for a solar photovoltaic thermal (PV/T) air collector, Appl. Energy,
975
2010; 87:2328–2339.
976 977
17. Riffat S.B. and Cuce E., A review on hybrid photovoltaic/thermal collectors and systems, Int. J. Low-Carbon Tech., 2011; doi: 10.1093/ijlct/ctr016; July 21.
978
18. Vokas G, Christandonis N, Skittides F, Hybrid photovoltaic thermal systems for
979
domestic heating and cooling-a theoretical approach, Solar Energy, 2006; 80: 607–15.
980
19. Makrides G, Zinsser B, Phinikarides A, Schubert, M. and Georghiou, G.E., Temperature
981
and thermal annealing effects on different photovoltaic technologies, Renew. Energy,
982
2012; 43: 407-17.
983
20. Lim JL, Woo SC, Jung TH, Min YK, Won CS, Ahn HK, Analysis of factor on the
984
temperature effect on the output of PV module, Trans. of the Korean Inst. of Elec.
985
Engineers, 2013; 62: 365-70.
986
21. Ronak D, Adnan I, Goh LJ, Ruslan MH, Kamaruzzaman S, Predicting the performance
987
of amorphous and crystalline silicon based photovoltaic solar thermal collectors, Energy
988
Conv. and Manag., 2011; 52: 1741- 47.
33
ACCEPTED MANUSCRIPT 989
22. Adiyabat, A., Kurokawa, K., Otani, K., Enebish, N., Batsukh, G., Battushig, M.,
990
Ochirvaani, D., Ganbat, B., Evaluation of solar energy potential and PV module
991
performance in the Gobi Desert of Mongolia, Prog. in Photovoltaics: Res. and Appl.,
992
2006; 14:553-66.
993
23. Skoplaki E, Palyvos JA, On the temperature dependence of photovoltaic module
994
electrical performance: A review of efficiency/power correlations, Solar Energy, 2009;
995
83: 614-24.
996
24. Nishioka K, Hatayama T, Uraoka Y, Fuyuki T, Hagihara R, Watanabe M, Field-test
997
analysis of PV system output characteristics focusing on module temperature, Solar
998
Energy Mat. and Solar Cells, 2003; 75: 665-71.
999 1000
25. Van Dyk, EE., Scott, BJ., Meyer, EL., Leitch, AWR., Temperature dependence of performance of crystalline silicon modules, S. Afr. J. Sci., 2000; 96: 198–200.
1001
26. Makrides G, Bastian Z, George EG, Markus S, Jurgen HW, Temperature behaviour of
1002
different photovoltaic systems installed in Cyprus and Germany, Solar Energy Mat. and
1003
Solar Cells, 2009; 93: 1095-99.
1004 1005 1006 1007 1008 1009
27. Kalogirou SA, Tripanagnostopoulos Y, Hybrid PV/T solar systems for domestic hot water and electricity production, Energy Conv. and Manag., 2006; 47: 3368–82. 28. Tiwari, A., Barnwal, P., Sandhu, G.S., Sodha, M.S., Energy metrics analysis of hybrid Photovoltaic (PV) modules, Appl. Energy, 2009; 86: 2615– 2625. 29. Tiwari, G.N., Mishra, R.K., Solanki, S.C., Photovoltaic modules and their applications: A review on thermal modeling, Appl. Energy, 2011; 88: 2287–2304.
1010
30. Tiwari, A., Sodha, M.S., Chanda, A., Joshi, J.C., Performance evaluation of photovoltaic
1011
thermal solar air collector for composite climate of India, Solar Energy Mat. & Solar
1012
Cells, 2006; 90: 175-189.
1013
31. Avireni S., Converging lens solar concentrator and their position controlled using a
1014
microprocessor for increasing the efficiency of solar photovoltaic energy conversion, 3rd
1015
Int. Conf.: Sci. of Elect., Techn. of Inf. and Tel., 2005; March 27-31: Tunisia.
1016 1017
32. Kerzmann, T., Schaefer, L., System simulation of a linear concentrating photovoltaic system with an active cooling system, Renew. Energy, 2012; 41, pp. 254-261.
34
ACCEPTED MANUSCRIPT 1018
33. El –Sayed, S.M., Anis, W.R., Hafez, I.M., The effect of temperature on the performance
1019
of PV array operating under concentration, Int. J. of Sci. & Tech. Res., 2015; 4, 8, pp.
1020
240-246.
1021
34. Xie, W.T., Dai, Y.J., Wang, R.Z., Sumathy, K., Concentrated solar energy applications
1022
using Fresnel lenses: A review, Renew. and Sust. Energy Rev., 2011; 15, pp. 2588- 2606.
1023
35. Tripanagnostopoulos, Y., Aspects and improvements of hybrid photovoltaic/thermal
1024 1025 1026
solar energy systems, Solar energy, 2007; 81: pp. 1117-1131. 36. Kalogirou, S.A., Use of TRNSYS for modelling and simulation of a hybrid pv-thermal solar system for Cyprus, Renewable Energy, 2001; 23: pp. 247-260.
1027
37. Herrando, M., Markides, C.N., Hellgardt, K., A UK-based assessment of hybrid PV and
1028
solar-thermal systems for domestic heating and power: System performance, Appl.
1029
Energy, 2014; 122: pp. 288-309.
1030
38. Laemmle M., Kroyer, T., Fortuin, S., Wiese, M., Hermann, M., Development and
1031
modelling of highly-efficient PVT collectors with low-emissivity coatings, Frauhofer
1032
Inst. for Solar Energy Systems ISE, 79110 Freiburg, Germany, 2016.
1033
39. Guarracino, I.,
Mellor, A., Ekins-Daukes, N.J., Markides, C.N., Dynamic coupled
1034
thermal-and-electrical modelling of sheet-and-tube hybrid photovoltaic/thermal (PVT)
1035
collectors, Appl. Thermal Engg., 2016; 101: 778-795.
1036
40. Ji, J., Han, J., Chow, T., Yi, H., Lu, J., He, W., Sun, W., Effect of fluid flow and packing
1037
factor on energy performance of a wall-mounted hybrid photovoltaic/water-heating
1038
collector system, Energy and Buildings, 2006; 38: pp. 1380-1387.
1039
41. Chow, T.T., Pei, G., Fong, K.F., Lin, Z., Chan, A.L.S., Ji, J., Energy and exergy analysis
1040
of photovoltaic-thermal collector with and without glass cover, Appl. Energy, 2009; 86:
1041
pp. 310-316.
1042
42. Nahar, A., Hasanuzzaman, M., Rahim, N.A., Numerical and experimental investigation
1043
on the performance of a photovoltaic thermal collector with parallel plate flow channel
1044
under different operating conditions in Malaysia, Solar Energy, 2017; 144: pp. 517–528.
1045
43. Sardarabadi, M., Passandideh-Fard, M., Experimental and numerical study of metal-
1046
oxides/water nanofluids as coolant in photovoltaic thermal systems (PVT), Solar Energy
1047
Mat. & Solar Cells, 2016; 15: pp. 7533–542
35
ACCEPTED MANUSCRIPT 1048 1049
44. Reddy, J.N. and Gartling, D.K., The Finite Element Method in Heat Transfer and Fluid Dynamics, CRC Press, Inc., Boca Raton, Florida, 1994.
1050
45. Nasrin, R. and Alim, M.A., Laminar free and forced magnetoconvection through an
1051
octagonal channel with a heat generating circular cylinder, J. of Naval Arch. and Marine
1052
Engg., 2012; 9(1): pp. 25-34.
1053
46. Chandrasekar M., Suresh S., Senthilkumar T., Karthikeyan MG, Passive cooling of
1054
standalone flat PV module with cotton wick structures, Energy Conv. and Manag., 2013;
1055
71: pp. 43-50.
1056
47. Bahaidarah H., Subhan A., Gandhidasan P., Rehman S., Performance evaluation of a PV
1057
(photovoltaic) module by back surface water cooling for hot climatic conditions, Energy,
1058
2013; 59: pp. 445-453.
1059
48. Bosanac, M., Sorensen, B., Katic, I., Sorensen, H., Nielsen, B., Badran, J.,
1060
Photovoltaic/Thermal solar collectors and their potential in Denmark, EFP project,
1061
1713/00-0014, Copenhagen, 21 May, 2003.
36