You are on page 1of 15

Case Studies in Thermal Engineering 28 (2021) 101361

Contents lists available at ScienceDirect

Case Studies in Thermal Engineering


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

Evaluation of nominal operating cell temperature (NOCT) of


glazed photovoltaic thermal module
Vat Sun a, Attakorn Asanakham b, c, Thoranis Deethayat b, c,
Tanongkiat Kiatsiriroat b, c, *
a
Energy Engineering Program, Department of Mechanical Engineering, Faculty of Engineering and Graduate School, Chiang Mai University, 239
Huay Kaew Road, Muang District, Chiang Mai, 50200, Thailand
b
Department of Mechanical Engineering, Faculty of Engineering, Chiang Mai University, 239 Huay Kaew Road, Muang District, Chiang Mai, 50200,
Thailand
c
Research Group for Renewable Energy, Faculty of Engineering, Chiang Mai University, 239 Huay Kaew Road, Muang District, Chiang Mai, 50200,
Thailand

A R T I C L E I N F O A B S T R A C T

Keywords: Nominal operating cell temperature (NOCT) was commonly used to evaluate photovoltaic (PV)
Glazed PVT module module temperature and this term is provided by the manufacturer but in case of photovoltaic
Performance thermal module, the NOCT was not fixed and varied with the circulating fluid mass flow rate and
Nominal operating cell temperature
its temperature. In this study, a novel approach on NOCT evaluation of glazed photovoltaic
Combined heat and power
thermal (PVTg) module for combined heat and power generation was presented by modifying the
method of solar collector testing for water heating. The thermal characteristics in terms of optical
and heat loss performances and finally the NOCT at various water flow rates and inlet temper­
atures could be evaluated. Experimental tests were performed with a PVTg module under Chiang
Mai climate, and it could be founded that the calculated module temperature, the outlet hot water
temperature and the generated electrical power agreed well with the experimental results.
Monthly performance of this module was also evaluated and compared with unglazed PVT and
conventional PV modules each having the same solar photovoltaic cell type. The average
generated electrical outputs were 28.07, 31.89, and 30.38 kWh/month, respectively, while the
glazed and the unglazed PVT modules could generate average thermal energy of 119.19 and
101.40 kWh/month, respectively.

1. Introduction

Among the renewable energy, photovoltaic (PV) has become a major contributor to global electricity generation, which shares
about 2.9% of global energy demand and saving 590 million tonnes of CO2 in 2018 [1]. As the PV module exposes to the sun, its module
temperature increases, and the power generation is reduced. The PV module conversion efficiency and power output were reduced by
0.4–0.5% per degree Celsius of the PV module temperature increment [2]. Therefore, several active and passive cooling techniques
have been performed to control the module temperature. Teo et al. [3] reduced PV module temperature by forced air cooling, and the

* Corresponding author. Department of Mechanical Engineering, Faculty of Engineering, Chiang Mai University, 239 Huay Kaew Road, Muang
District, Chiang Mai, 50200, Thailand.
E-mail address: tanongkiat_k@yahoo.com (T. Kiatsiriroat).

https://doi.org/10.1016/j.csite.2021.101361
Received 5 November 2020; Received in revised form 13 August 2021; Accepted 16 August 2021
Available online 21 August 2021
2214-157X/© 2021 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
V. Sun et al. Case Studies in Thermal Engineering 28 (2021) 101361

module efficiency could be increased from 8–9% to 12–14%. Abdolzadeh and Ameri [4] cooled down PV module temperature by water
spray over the front of PV module, and the mean module efficiency was increased by 3.26%. Sun et al. [5] used passive conductive
cooling with phase change material installed at the back of PV module, and the average electrical power output could be increased by
4.3%. The active cooling methods seems to be preferable when the extracted heat from the module could be used for other thermal
applications in terms of hot air or hot water. With combined heat and power generation, PV module is named Photovoltaic Thermal
(PVT) module and the overall efficiency of the PVT module might reach 70%, with electrical efficiency between 15 and 20% and the
thermal efficiency of 50% [6].
Electrical power generated by PV module could be predicted from its nominal operating cell temperature (NOCT) [5,7,8] which is
normally a constant provided by manufacturer. This term is used to evaluate the average module temperature which was used to
calculate the electrical power output. However, for PVT module, the NOCT is not available since this term varies with flow rate and
inlet temperature of cooling fluid [9]. Then many approaches have been developed to estimate the PVT module temperature. Guar­
racino et al. [10,11] and Rejeb et al. [12] used three-dimensional transient methods and finite volume method with energy balance
equations for six main components: transparent cover, PV module, plate absorber, water tube, circulating fluid, and insulation
including the PVT module thermal properties and the related heat transfer coefficients to evaluate the PVT module temperature and
the electrical power output. Some approaches with steady-state models were developed by Sahota and Tiwari [13], Tiwari et al. [14],
and Atheaya et al. [15]. The heat transfer from the solar cell module to the surrounding ambient and that from the solar cell module to
the absorber plate were evaluated including the heat transfer from the absorber to the circulating fluid and then the PVT module
temperature could be estimated. However, the calculations were very complicated since a lot of thermal properties and related heat
transfer coefficients were needed. A computational fluid dynamics software such as ANSYS Fluent or COMSOL was used by Arslan et al.
[16], Ramdani and Ould-Lahoucine [17], Misha et al. [18], Nahar et al. [19] to predict module temperature, water outlet temperature,
and thermal performances of the PVT module. Numerical simulations could be carried out from the governing equations such as
continuity, momentum, and energy equations in a set of finite volumes. The surface temperatures of the PVT module were estimated at
a given ambient temperature and some prescribed solar radiation levels and flow rates. The simulation was time consuming, and it was
very complicated and not suitable when the input weather data were not constant during the daytime.
A simplified method on NOCT evaluation of an unglazed PVT module for electrical power and hot water generation was developed
by Sun et al. [9]. The NOCT could be evaluated from the PVT module optical efficiency, (τα), and overall heat loss coefficient, UL , of
which the values could be experimentally investigated following the method of solar collector testing for liquid heating when the
module temperature was recorded [20]. The NOCT was found to depend on the inlet temperature and the mass flow rate of the working
fluid. With the evaluated NOCT, the module temperature, the generated electrical power and the hot water temperature could be
estimated, and the results agreed very well with the measured data.
In case of glazed PVT (PVTg) module, the evaluation of the NOCT is more complicated since the module temperature is rather
difficult to be measured. Therefore, in this study a set of models following the method of solar thermal collector for water heating
including the PVT panel configuration was developed to find out the PVT module thermal characteristics such as F , FR , FR (τα), and

FR UL [20] and these values were used to find out the NOCT at any mass flow rate and inlet temperature which was the aim of this study.
The NOCT could also be used to evaluate the PVTg module temperature, the generated hot water temperature, and the generated
electrical power. Experimental tests were also performed to verify the model. In addition, monthly performances on heat and power
generation of the PVTg module were also evaluated and compared with the unglazed PVT module and typical PV module each having
the same solar photovoltaic cell type.

2. Methodology

2.1. PVT module performance

For calculating the maximum electrical power output of PV or PVT module (Pm ), the following linear model is given by Refs. [5,
21–23]:
τ g · IT
Pm = Pm,stc · (1 − γ · (TPV − 25)) · (1)
1000

where τg is the glass transmittance of PVT module with glass cover (τg =1 for unglazed PVT module and τg = 0.85 for glazed PVT
module [24]), Pm,stc is the maximum electrical power at Standard Testing Condition when the solar irradiance is at 1000 W/m2 and the
PV or the PVT module temperature is at 25 ◦ C, γ is the temperature coefficient of maximum power and IT is the solar irradiance on tilted
plane. TPV is the PV or PVT module temperature which could be calculated by Refs. [20,25]:
IT
TPV = Ta + (NOCT − 20) · (2)
800

where Ta is the ambient temperature, NOCT is the module temperature at the ambient temperature of 20 ◦ C and the solar irradiance of
800 W/m2. The value is approximately 45 ± 2 ◦ C for monocrystalline and polycrystalline PV modules. Generally, the NOCT is constant
for PV module and the value is given by manufacturer but in case of the PVT, the NOCT is not defined since it depends on the fluid mass
flow rate (ṁ) and the inlet fluid temperature (Tfi ) [9].
Following the solar collector equation for fluid heating,

2
V. Sun et al. Case Studies in Thermal Engineering 28 (2021) 101361

Fig. 1. Flow chart for calculating PVT module performance by NOCT method.

TPV − Ta
ηth = (τα) − UL , (3)
IT

or
Tfm − Ta
(4)
′ ′
ηth = F (τα) − (F UL ) ,
IT

and
Tfi − Ta
ηth = FR (τα) − (FR UL ) , (5)
IT

where ηth is the thermal efficiency, (τα) is the optical efficiency, UL is the overall heat loss coefficient, Tfi is the inlet fluid temperature,
Tfm is the mean fluid temperature, FR is the module heat removal factor, and F is the module efficiency factor. F (τα), F UL , FR (τα), and
′ ′ ′

FR UL are thermal characteristics those could be found from the thermal performance followed the ASHRAE standard for solar collector
testing [26].

3
V. Sun et al. Case Studies in Thermal Engineering 28 (2021) 101361

Fig. 2. Schematic sketch of the experimental setup.

By Eqs. (2), (3) and (5)


(NOCT − 20) TPV − Ta Tfi − Ta FR (τα)
= = FR · + · (1 − FR ) (6)
800 IT IT FR UL

then
Tfi − Ta FR (τα)
NOCT = 800FR · + · (1 − FR )800 + 20 (7)
IT FR U L
When the mass flow rate becomes zero, the FR value is also zero, then from Eq. (7): NOCT = (τα)/UL · 800 + 20. At this condition the
PVT module will act as a PV module [20,27].
FR could be calculated from [20].

4
V. Sun et al. Case Studies in Thermal Engineering 28 (2021) 101361

Table 1
Specifications of PVTg, PVT and PV modules.
Cell type PVTg PVT [9] PV reference

Mono-crystalline Mono-crystalline Mono-crystalline

Dimensions 870 mm × 1640 mm × 105 mm 828 mm × 1601 mm × 90 mm –


Aperture area 1.427 m2 1.326 m2 1.326 m2
Glass cover Yes No No
Weight 34.4 kg 24.4 kg –
Number of cells 72 72 72
Cell dimensions 125 mm × 125 mm 125 mm × 125 mm 125 mm × 125 mm
Maximum power, Pm 200 Wp 200 Wp 200 Wp
Electrical efficiency of module, ηstc 14.02% 15.08% 15.08%
Temperature coefficients of Pm , γ -0.45%/◦ C -0.45%/◦ C -0.45%/◦ C
NOCT Not available Not available 45 ◦ C ± 2 ◦ C

FR ṁcp [ ( /( ))]
(8)

= 1 − exp − Ac F UL ṁcp
F ′ Ac F ′ U L

In case of fin-tube absorber plate as in this study (Fig. A1 in the Appendix), F could be calculated from

1/UL
(9)

F = [ ]
1
W UL (D+(W− D)F)
+ c1b + πD1i hfi

or
[ ]
1 D + (W − D)F 1 W W
= ′ − − (10)
UL W F UL cb πDi hfi

where D is the tube diameter, W is the tube spacing, F is the fin efficiency, cb is the bond conductance, Di is the tube diameter (inside),
hfi is the heat transfer coefficient inside tubes, and for laminar flow, ReD < 2300, the value could be determined from [10,28].
kf
hfi = 4.36 (11)
Di
The fin efficiency could be calculated by
√̅̅̅̅̅̅
tanh[m(W − D)/2] UL
F= ,m= (12)
m(W − D)/2 kδ

where k is the plate thermal conductivity and δ is the plate thickness. A trial-and-error method was deployed to estimate the UL of the
PVT module in Eq. (10). Fig. A2 in the Appendix showed the calculated UL of the tested PVT module which was 6.6382 W/(m2⋅K) at the
water flow rate of 1.72 LPM. With a given value of mass flow rate, hfi , UL followed by F and FR could be calculated. The NOCT could be

calculated from Eq. (7).


When the mass flow rate is different from the standard test value, the FR UL and FR (τα) could be corrected as [20].

ṁcp [ ( / )]⃒
1 − exp − A F

U ṁc ⃒

A c F UL c L p ⃒
r= ⃒use . (13)
ṁcp [ ( / )]⃒
′ 1 − exp − A c F ′
U L ṁc p ⃒

A c F UL
ref

Then the FR (τα) and FR UL of the use condition could be corrected by


FR UL |use FR (τα)|use
r= = (14)
FR UL |ref FR (τα)|ref

The useful heat gain from PVT module could be calculated by


[ ( )]
Q̇u = Ac FR (τα)IT − FR UL Tfi − Ta (15)

and the fluid outlet temperature, Tfo could be calculated by


Ac [ ( )]
Tfo = Tfi + FR (τα)IT − FR UL Tfi − Ta . (16)
ṁcp

Fig. 1 shows the calculation procedure of the NOCT method. The input data of the calculation covered FR (τα)|ref and FR UL |ref which

5
V. Sun et al. Case Studies in Thermal Engineering 28 (2021) 101361

Fig. 3. Thermal performance of the tested PVTg module following the ASHRAE Standard 93–2003 when the water mass flow rate was 1.72 LPM. (a)
thermal efficiency vs. (Tfi − Ta )/IT and (b) thermal efficiency vs. (Tfm − Ta )/IT . The raw data came from the manufacturer [29] given in Table A1 of
the Appendix.

Table 2
Thermal characteristics obtained from Fig. 3 and the calculated of UL , F , and FR at the reference water flow rate of 1.72 LPM.

Experimental results from manufacturer raw data Calculated results


2 2
ṁ LPM ṁcp W/K Ac m FR (τα) FR UL W/(m ⋅K) F (τα)
′ 2
F UL W/(m ⋅K)

FR (τα)/FR UL UL W/(m2⋅K) F

FR
1.72 118.43 1.427 0. 4667 5.424 0.4822 5.6014 0.0861 6.6382 0.8446 0.8171

are reference values obtained from experiment at a reference mass flow rate ṁ|ref and the values of the solar radiation level, IT , and the
ambient temperature Ta . At any input mass flow rate ṁ with a given inlet water temperature Tfi , the output parameters which are
module temperature TPV , generated power Pm , outlet hot water temperature Tfo and useful heat gain Q̇u , could be evaluated. The
calculated values were compared with the experiment data to verify the correction of the model.

6
V. Sun et al. Case Studies in Thermal Engineering 28 (2021) 101361

Table 3
Thermal characteristics and other parameters at different water mass flow rates.
ṁ LPM ṁcp W/K Ac m2 FR FR (τα) FR UL W/(m2⋅K) (τα) 800FR W/m2 FR (τα)
· (1 − FR )800 + 20 ◦ C
FR UL
0 0 1.427 0 0 0 0.5714 0 88.83
0.5 34.833 1.427 0.7547 0.4311 5.0101 0.5714 603.79 36.88
1 69.67 1.427 0.7980 0.4558 5.2972 0.5714 638.39 33.91
1.72 118.433 1.427 0.8171 0.4667 5.4240 0.5714 653.67 32.59
2.4 167.2 1.427 0.8248 0.4711 5.4750 0.5714 659.82 32.06
4 278.667 1.427 0.8326 0.4756 5.5273 0.5714 666.12 31.52
6 418 1.427 0.8366 0.4779 5.5536 0.5714 669.30 31.25

Fig. 4. Calculated NOCT of PVTg module: (a) with water mass flow rate of 0–6 LPM and (b) compared with unglazed PVT module [9]. The inlet
water temperature, Tfi was 20–85 ◦ C.

7
V. Sun et al. Case Studies in Thermal Engineering 28 (2021) 101361

Fig. 5. Validation of PVTg module temperature, water outlet temperature and maximum power output on some clear sky days: (a) March 16, 2018
and (b) April 5, 2018 (the water flow rate of 1 LPM).

2.2. Experimental setup

A tested glazed PVT module (PVTg) having a maximum electrical power of 200 Wp with an aperture area of 1.427 m2 was con­
nected with a well-insulated 100-L water tank, as shown in Fig. 2. The specifications of glazed PVT (PVTg) including the unglazed PVT
(PVT) and the typical PV modules each having the same solar cell type were given in Table 1. The experiment was done outdoor under
Chiang Mai climate, and the PVT module was facing south with an inclined plane of 18◦ . A set of K-type thermocouples were used to
measure the inlet, the outlet, the module and the ambient temperatures of the PVT module, and the data was logged by Huato S220-T8.
The solar irradiance was measured by a pyranometer CMP3 installed on the inclined plane of the PVT module monitored a datalogger
EX9018. The solar analyzer was used to measure the maximum electrical power. The data were recorded every 1 min except the power
analyzer of which the data were recorded every 5 min. The experiment was performed between 9:00 a.m. to 4:00 p.m.

8
V. Sun et al. Case Studies in Thermal Engineering 28 (2021) 101361

Fig. 6. Validation of PVT module temperature and its maximum electrical power output on May 2, 2019, which is a cloudy day: (a) the measured
solar irradiance, the ambient temperature, and the water inlet temperature, (b) the PVT module temperature, and (c) the maximum electrical power
output of PVT module.

3. Results and discussion

3.1. Experimental results

Fig. 3 demonstrated thermal performance of the tested commercial glazed PVT module based on Eqs. (4) and (5) in a form of ηth vs.
(Tfi − Ta )/IT diagram following the ASHRAE Standard 93–2003 [26] for solar collector testing. The raw data given in table A1 (in the
Appendix) were obtained from the manufacturer test report [29] at water flow rate of 1.72 LPM. The relation between ηth and
(Tfi − Ta )/IT was in a linear form whereas the FR (τα) and FR UL were the intercept on the ηth axis and the slope of the relation,
respectively. Similarly, F (τα) and F UL based on Eq. (4) were the intercept on the ηth axis and the slope of the relation in ηth vs.
′ ′

(Tfm − Ta )/IT diagram. The tests were done with the solar radiation in a range of 890–1010 W/m2 and the ambient temperature in a
range of 22.2–28 ◦ C.
The thermal characteristic parameters FR (τα) FR UL , F (τα) and F UL and the calculation of UL , F and FR from Eqs. 8–12 at the
′ ′ ′

reference water flow rate of 1.72 LPM were also given in Table 2.

9
V. Sun et al. Case Studies in Thermal Engineering 28 (2021) 101361

Fig. 7. Chiang Mai weather data: the solar irradiance on the 18-degree tilted surface (IT ) and the ambient temperature (Ta ). The time starts at 0:00
a.m. of January 1st.

3.2. NOCT evaluation

The water flow rate at 1.72 LPM was taken as a reference then with other flow rates, the thermal characteristics and others related
parameters was changing and those could determine from Eqs. (13) and (14). The results were shown in Table 3 when the flow rate was
in a range of 0–6 LPM.
Fig. 4 shows the relation of NOCT calculated from Eq. (7) and (Tfi − Ta )/IT at a given mass flow rate. The intercept on the NOCT axis
was the value of FR (τα)/FR UL · (1 − FR )800 + 20 and 800FR was the slope of the performance curve. These values could be taken from
Table 3.
From Fig. 4(a), it could be seen that the NOCT increased with decreasing the fluid flow rate, but when (Tfi − Ta )/IT was over 0.06,
the NOCT at any value of (Tfi − Ta )/IT was nearly steady for all flow rates. It could be noted that when the (Tfi − Ta )/IT was over 0.06,
the inlet water was rather high (might be up to 80 ◦ C) and it will cause a problem to the PV module. In the case of zero water flow rate,
the module acted like a typical PV, and its NOCT was constant like a typical PV. However, its value (88.83 ◦ C) was rather high
compared with those of the typical PV (45 ◦ C [20]) and the unglazed PVT type (61 ◦ C from Sun et al. data [9]). Increase of the inlet
temperature also resulted in increase of the NOCT.
Fig. 4(b) shows comparison of the NOCT - (Tfi − Ta )/IT relations of the glazed and unglazed PVT modules at water flow rates of 1 and
6 LPM. It could be seen that the slope which was 800FR or FR of the glazed PVT module was higher. The FR also affected the NOCT in Eq.
(7). At low (Tfi − Ta )/IT , for example Tfi ≈ Ta , the term (1 − FR ) in Eq. (7) dominated the NOCT value and due to the higher FR for the
glazed PVT module the NOCT of the glazed PVT module was lower than that of unglazed PVT type. At high (Tfi − Ta )/ IT , the first term
of the right-hand side in Eq. (7) controlled the NOCT then in this case the NOCT of the glazed PVT module was higher. It could be noted
that higher the mass flow rate resulted in lower the NOCT value.

3.3. Validation of NOCT model

The experiment was done in Chiang Mai on March 16, 2018 and (b) April 5, 2018 which are clear sky days as shown in Fig. 5. The

10
V. Sun et al. Case Studies in Thermal Engineering 28 (2021) 101361

Fig. 8. Simulated results on monthly electrical and thermal energies generated by PV, PVT, PVTg under Chiang Mai’s weather, the water flow rate
of 1 LPM, the inlet water temperature was 25 ◦ C.

PVTg module was tested at the water flow rate of 1 LPM and with the values of FR , FR (τα), FR UL , Ac and ṁcp given in Table 3, the NOCT,
the PVTg module temperature, the outlet water temperature and the generated maximum electrical power could be calculated by Eqs.
(1), (2), (7) and (16), respectively. The measured weather data (IT and Ta ) and the inlet temperature (Tfi ) were the input values for the
calculation. It was noticed that the calculated results agreed very well with the experimental data. The mean absolute percentage error
(MAPE) for the PVTg module temperature, the outlet temperature, and the maximum power output were less than 3.26%, respectively.
Fig. 6 also shows the PVTg module performance on a cloudy day. It could be seen that the simulated results agreed very well with
the measured data. For the generated maximum power, the measurement was terminated after around 13:00 o’clock due to rain
showers. The MAPE of PVTg module temperature was 2.91% and the MAPE of maximum electrical output was 4.45%.

3.4. Monthly performance

The monthly performance of the glazed PVT (PVTg) module was also simulated under Chiang Mai weather data and compared with
those of the unglazed (PVT) module and the typical PV module having the same solar cell type.
The typical meteorological data included the global horizontal radiation, the diffuse horizontal radiation, and the ambient

11
V. Sun et al. Case Studies in Thermal Engineering 28 (2021) 101361

temperature of Chiang Mai were obtained from the Photovoltaic Geographical Information System (PVGIS) under European Com­
mission’s Science and Knowledge Service [30], and the calculated total solar irradiance on the 18-degree tilted surface was given in
Fig. 7. The simulations of the PVTg and PVT module performances were done with the water flow rate of 1 LPM and the inlet water
temperature at 25 ◦ C.
Fig. 8 shows the monthly electrical and thermal energy generated by the PVTg, the PVT, and the typical PV modules. The NOCT of
the glazed and the unglazed modules were calculated by Eq. (7), and the NOCT of the typical PV module was 45 ◦ C. The PVTg, the PVT
and the typical PV modules could generate average electrical energy output of 28.07 kWh/month, 31.89 kWh/month and 30.38 kWh/
month, respectively (Fig. 8(a)). It meant the PVTg generated electrical energy 7.62% less than that of the typical PV module even the
module was cooled by water, but the PVT module generated 4.94% more than that of the typical PV module. For the thermal energy,
the PVTg and PVT module could generate thermal energy of 119.19 kWh/month and 101.40 kWh/month, which meant the PVTg
module could generate thermal energy 17.54% higher than that of the PVT module. In conclusion, the glazed PVT module was rec­
ommended when high thermal energy was needed, but the unglazed PVT module was preferable for electrical energy.

4. Conclusions

A novel approach on evaluation of NOCT of glazed PVT module for combined heat and power generation was presented. By
modifying the testing method of solar collector for water heating, the thermal characteristics such as FR , FR (τα), and FR UL at any water
flow rate and its temperature could be estimated and finally the NOCT. With the calculated NOCT, the water outlet temperature, the
module temperature and the electrical power output could be evaluated, and the results agreed very well with those from the outdoor
experiments, of which the mean absolute percentage error was not over 3.26% on clear sky day and 4.45% on cloudy day under Chiang
Mai weather data. Moreover, monthly performance of the glazed and the unglazed PVT modules were also evaluated and compared
with the typical PV module each having the same photovoltaic cell type. The glazed PVT module was preferable when high thermal
energy was needed compared with the unglazed PVT module and vice versa for high electrical energy requirement. The monthly
average of the generated electrical outputs were 28.07, 31.89, and 30.38 kWh/month, respectively, while the glazed and unglazed
modules showed the average thermal energy of 119.19 and 101.40 kWh/month, respectively.

Authorship contributions

Please indicate the specific contributions made by each author (list the authors’ initials followed by their surnames, e.g., Y.L.
Cheung). The name of each author must appear at least once in each of the three categories below. Conception and design of study: V.
Sun, A. Asanakham, T. Deethayat, T. Kiatsiriroat, Acquisition of data: V. Sun, A. Asanakham, Analysis and/or interpretation of data: V.
Sun, A. Asanakham, T. Deethayat, T. Kiatsiriroat, Drafting the manuscript: V. Sun, A. Asanakham, T. Deethayat, T. Kiatsiriroat,
Revising the manuscript critically for important intellectual content: V. Sun, A. Asanakham, T. Deethayat, T. Kiatsiriroat, Approval of
the version of the manuscript to be published (the names of all authors must be listed): V. Sun, A. Asanakham, T. Deethayat, T.
Kiatsiriroat.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Acknowledgements

This research project is supported by the Department of Mechanical Engineering (Research Assistant Program), Faculty of Engi­
neering and Research Group for Renewable Energy, Chiang Mai University, and the National Research Council of Thailand under the
Project on Formation of Sustainable Green Communities by Alternative Energy.

Nomenclature

Abbreviation
LPM liter per minute
NOCT nominal operating cell temperature, ◦ C
PV photovoltaic
PVT photovoltaic thermal

Symbol
Ac area of PVT module, m2
cp specific heat capacity, J/(kg⋅K)
cb bond conductance
D tube diameter, m

12
V. Sun et al. Case Studies in Thermal Engineering 28 (2021) 101361

Di internal tube diameter, m


F module efficiency factor

FR module heat removal factor


hfi fluid convection
IT solar irradiance, W/m2
kf fluid thermal conductivity, W/(m⋅K)
k plate thermal conductivity, W/(m⋅K)
ṁ mass flow rate, kg/s
Pm maximum power point, W
Pm,stc maximum power point at standard testing condition, W
Q̇u useful heat, W
Ta ambient temperature, ◦ C
Tfi inlet temperature, ◦ C
Tfm mean fluid temperature, ◦ C
Tfo outlet temperature, ◦ C
TPV PVT module temperature, ◦ C
UL overall heat loss coefficient, W/(m2⋅K)
W tube spacing, m
w wind speed, m/s
(τα) module optical efficiency
γ temperature coefficient of Pm , %/◦ C
ηth thermal efficiency
τg glass transmittance
δ plate thickness, m

Appendix

Fig. A.1. Configuration of the tested glazed PVTg and channel flow structure [31].

13
V. Sun et al. Case Studies in Thermal Engineering 28 (2021) 101361

Fig. A.2. Calculated UL by trial-and-error of the tested PVTg module. The water mass flow rate was 1.72 LPM.
Table A.1
Tested results from manufacturer [29].

Number Ta ◦ C IT W/m2 wm/s Tfi ◦ C Tfo ◦ C ṁLPM Tfm ◦ C

1 22.2 983 1.63 15 20.7 1.72 17.85


2 22.4 977 1.57 15.08 21.02 1.72 18.05
3 22.3 951 1.96 15.09 20.6 1.73 17.85
4 22.6 944 1.99 15.12 20.8 1.73 17.96
5 23.8 918 2.09 26 31.22 1.71 28.61
6 24.2 934 2.18 26.1 31.2 1.73 28.65
7 24.8 957 2.23 26.28 31.64 1.71 28.96
8 23.9 893 1.57 26.05 30.91 1.72 28.48
9 23.1 1006 1.51 38.81 43.44 1.72 41.12
10 22.8 976 1.41 37.96 42.49 1.73 40.23
11 22.7 893 1.78 37.4 41.5 1.72 39.4
12 23.3 925 1.58 38.18 42.31 1.74 40.25
13 27.4 964 1.33 53.03 56.79 1.72 54.91
14 26 948 2.31 51.69 55.19 1.72 53.44
15 26.8 973 2.36 52.96 56.55 1.73 54.76
16 28 992 1.1 54.21 57.86 1.73 56.04
Mean absolute percentage error.

n ⃒ ⃒
100% ∑ ⃒y i − x i ⃒
MAPE = ⃒
⃒ xi ⃒
⃒ (A.1)
n i=1

where yi is the calculated value, xi is the measured value, and n is the number of experimental data.

References

[1] G. Masson, I. Kaizuka, Photovoltaic power systems programme, 2019. https://iea-pvps.org/wp-content/uploads/2020/02/5319-iea-pvps-report-2019-08-lr.pdf.


[2] V. Sun, A. Asanakham, T. Deethayat, T. Kiatsiriroat, Study on phase change material and its appropriate thickness for controlling solar cell module temperature,
Int. J. Ambient Energy (2018) 1–10, https://doi.org/10.1080/01430750.2018.1443500.
[3] H.G. Teo, P.S. Lee, M.N.A. Hawlader, An active cooling system for photovoltaic modules, Appl. Energy 90 (2012) 309–315, https://doi.org/10.1016/j.
apenergy.2011.01.017.
[4] M. Abdolzadeh, M. Ameri, Improving the effectiveness of a photovoltaic water pumping system by spraying water over the front of photovoltaic cells, Renew.
Energy 34 (2009) 91–96, https://doi.org/10.1016/j.renene.2008.03.024.
[5] V. Sun, A. Asanakham, T. Deethayat, T. Kiatsiriroat, Increase of power generation from solar cell module by controlling its module temperature with phase
change material, J. Mech. Sci. Technol. 34 (2020) 2609–2618, https://doi.org/10.1007/s12206-020-0336-8.
[6] A. Ramos, M.A. Chatzopoulou, I. Guarracino, J. Freeman, C.N. Markides, Hybrid photovoltaic-thermal solar systems for combined heating, cooling and power
provision in the urban environment, Energy Convers. Manag. (2017), https://doi.org/10.1016/j.enconman.2017.03.024.
[7] E. Skoplaki, J.A. Palyvos, On the temperature dependence of photovoltaic module electrical performance: a review of efficiency/power correlations, Sol. Energy
83 (2009) 614–624, https://doi.org/10.1016/j.solener.2008.10.008.
[8] R.G. Ross Jr., Flat-plate photovoltaic array design optimization, in: 14th Photovolt. Spec. Conf., 1980, pp. 1126–1132.
[9] V. Sun, A. Asanakham, T. Deethayat, T. Kiatsiriroat, A new method for evaluating nominal operating cell temperature (NOCT) of unglazed photovoltaic thermal
module, Energy Rep. 6 (2020) 1029–1042, https://doi.org/10.1016/j.egyr.2020.04.026.
[10] I. Guarracino, A. Mellor, N.J. Ekins-Daukes, C.N. Markides, Dynamic coupled thermal-and-electrical modelling of sheet-and-tube hybrid photovoltaic/thermal
(PVT) collectors, Appl. Therm. Eng. 101 (2016) 778–795, https://doi.org/10.1016/J.APPLTHERMALENG.2016.02.056.

14
V. Sun et al. Case Studies in Thermal Engineering 28 (2021) 101361

[11] I. Guarracino, J. Freeman, A. Ramos, S.A. Kalogirou, N.J. Ekins-Daukes, C.N. Markides, Systematic testing of hybrid PV-thermal (PVT) solar collectors in steady-
state and dynamic outdoor conditions, Appl. Energy 240 (2019) 1014–1030, https://doi.org/10.1016/J.APENERGY.2018.12.049.
[12] O. Rejeb, H. Dhaou, A. Jemni, A numerical investigation of a photovoltaic thermal (PV/T) collector, Renew. Energy 77 (2015) 43–50, https://doi.org/10.1016/
j.renene.2014.12.012.
[13] L. Sahota, G.N. Tiwari, Review on series connected photovoltaic thermal (PVT) systems: analytical and experimental studies, Sol. Energy 150 (2017) 96–127,
https://doi.org/10.1016/j.solener.2017.04.023.
[14] G.N. Tiwari, R.K. Mishra, S.C. Solanki, Photovoltaic modules and their applications: a review on thermal modelling, Appl. Energy 88 (2011) 2287–2304,
https://doi.org/10.1016/J.APENERGY.2011.01.005.
[15] D. Atheaya, A. Tiwari, G.N. Tiwari, I.M. Al-Helal, Performance evaluation of inverted absorber photovoltaic thermal compound parabolic concentrator (PVT-
CPC): constant flow rate mode, Appl. Energy 167 (2016) 70–79, https://doi.org/10.1016/J.APENERGY.2016.01.023.
[16] E. Arslan, M. Aktaş, Ö.F. Can, Experimental and numerical investigation of a novel photovoltaic thermal (PV/T) collector with the energy and exergy analysis,
J. Clean. Prod. 276 (2020), 123255, https://doi.org/10.1016/j.jclepro.2020.123255.
[17] H. Ramdani, C. Ould-Lahoucine, Study on the overall energy and exergy performances of a novel water-based hybrid photovoltaic-thermal solar collector,
Energy Convers. Manag. 222 (2020), 113238, https://doi.org/10.1016/j.enconman.2020.113238.
[18] S. Misha, A.L. Abdullah, N. Tamaldin, M.A.M. Rosli, F.A. Sachit, Simulation CFD and experimental investigation of PVT water system under natural Malaysian
weather conditions, Energy Rep. (2019), https://doi.org/10.1016/j.egyr.2019.11.162.
[19] A. Nahar, M. Hasanuzzaman, N.A. Rahim, Numerical and experimental investigation on the performance of a photovoltaic thermal collector with parallel plate
flow channel under different operating conditions in Malaysia, Sol. Energy 144 (2017) 517–528, https://doi.org/10.1016/j.solener.2017.01.041.
[20] J.A. Duffie, W.A. Beckman, Solar Engineering of Thermal Processes, Wiley, 2013, in: https://www.wiley.com/en-us/Solar+Engineering+of+Thermal+
Processes%2C+4th+Edition-p-9780470873663. (Accessed 8 April 2019).
[21] L.S. Pantic, T.M. Pavlović, D.D. Milosavljević, I.S. Radonjic, M.K. Radovic, G. Sazhko, The assessment of different models to predict solar module temperature,
output power and efficiency for Nis, Serbia, Energy 109 (2016) 38–48, https://doi.org/10.1016/j.energy.2016.04.090.
[22] E. Yandri, Development and experiment on the performance of polymeric hybrid Photovoltaic Thermal (PVT) collector with halogen solar simulator, Sol. Energy
Mater. Sol. Cells 201 (2019), 110066, https://doi.org/10.1016/j.solmat.2019.110066.
[23] J. Song, B. Sobhani, Energy and exergy performance of an integrated desiccant cooling system with photovoltaic/thermal using phase change material and
maisotsenko cooler, J. Energy Storage. 32 (2020), 101698, https://doi.org/10.1016/j.est.2020.101698.
[24] F. Giovannetti, S. Föste, N. Ehrmann, G. Rockendorf, High transmittance, low emissivity glass covers for flat plate collectors: applications and performance, in:
Energy Procedia, Elsevier Ltd, 2012, pp. 106–115, https://doi.org/10.1016/j.egypro.2012.11.014.
[25] R.G.J. Ross, R. G Jr., Flat-plate Photovoltaic Array Design Optimization, Photovolt. Spec. Conf. 14th, San Diego, Calif., January 7-10, 1980, Conf. Rec. (A81-
27076 11-44) New York, Inst. Electr. Electron. Eng. Inc., 1980, pp. 1126–1132 (1980) 1126–1132, http://adsabs.harvard.edu/abs/1980pvsp.conf.1126R.
(Accessed 17 September 2019).
[26] A. Standard, Standard 93-2003 “Methods of Testing to Determine the Performance of Solar Collectors, ASHRAE, Atlanta, 2003.
[27] Y. Zhang, T. Ma, P. Elia Campana, Y. Yamaguchi, Y. Dai, A techno-economic sizing method for grid-connected household photovoltaic battery systems, Appl.
Energy 269 (2020), 115106, https://doi.org/10.1016/j.apenergy.2020.115106.
[28] F.P. Incropera, D.P. Dewitt, T.L. Bergman, A.S. Lavine, Fundamentals of Heat and Mass Transfer, 1985, John Wiley, Hoboken, NJ, 2002, pp. 939–940.
[29] Test Report for Solimpeks Volther POWERTHERM According to EN 12975-2: 2006: Thermal Solar Systems and Components - Solar Collectors – Part 2: Test
Methods, Rivalta Scrivia, 2011. http://intergeo.sk/wp-content/uploads/GLAZED-TR-UNIEN12975-2-ENG-ITA-COMPLETO-1.pdf.
[30] Photovoltaic Geographical Information System (PVGIS) | EU Science Hub, (n.d.). https://ec.europa.eu/jrc/en/pvgis (accessed October 8, 2020).
[31] Powervolt | solimpeks solar corp, 2016. http://www.solimpeks.com/product/volther-powervolt/. (Accessed 29 October 2018).

15

You might also like