Mattei 2006

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Renewable Energy 31 (2006) 553–567

www.elsevier.com/locate/renene

Technical note

Calculation of the polycrystalline PV module


temperature using a simple method of energy balance
M. Mattei, G. Notton*, C. Cristofari, M. Muselli, P. Poggi
Laboratoire Systèmes Physiques de l 0 Environnement, Université de Corse Pascal Paoli,
UMR CNRS 6134, Route des Sanguinaires, F-20000 Ajaccio, France
Received 9 July 2004; accepted 7 March 2005
Available online 13 May 2005

Abstract
The performance of a photovoltaic module is studied versus environmental variables such as solar
irradiance, ambient temperature and wind speed. Two types of simplified models are studied in this
paper: a PV module temperature model and a PV module electrical efficiency model. These models
have been validated utilizing experimental data from two experiments: a 850 Wp grid connected
photovoltaic system and a p-Si module with eight temperature sensors integrated into the module.
Both models have been coupled to determine the PV array output power versus the three
meteorological parameters. This simple model using a simple energy balance and neglecting the
radiation effects is in good agreement with the experimental data.
q 2005 Elsevier Ltd. All rights reserved.

Keywords: Energy balance; Photovoltaic module; Temperature estimation

1. Introduction

It is well-known that most of the solar radiation absorbed by a photovoltaic (PV) panel
is not converted to electricity but contributes to increase the temperature of the module,
thus reducing the electrical efficiency.
This fact leads many researchers to develop hybrid PV/thermal collectors (PV/T) which
generate electric power and simultaneously produce hot water [1–3] or hot air [3,4]. The
photovoltaic cells are in thermal contact with a solar heat absorber and the excess heat

* Corresponding author. Tel.: C33 4 9552 4152; fax: C33 4 9552 4142.
E-mail address: gilles.notton@univ-corse.fr (G. Notton).

0960-1481/$ - see front matter q 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.renene.2005.03.010
554 M. Mattei et al. / Renewable Energy 31 (2006) 553–567

generated by the photovoltaic cells serves as an input for the thermal system. During the
operation, a heat carrier fluid removes heat from the absorber and PV cells. These cooled
cells then operate at a low and stable temperature and their electrical production increases
because their efficiency is decreasing with the temperature. The collected heat can be used
as preheated water or air.
The first step consists in knowing the behaviour of the PV module alone without the
absorber, the glass and the insulation. We want to study the influence of meteorological
parameters on the PV temperature and on the electrical efficiency.

2. Temperature influence on PV module performances

2.1. Some consideration and literature review

Two important parameters of the I–V curve for a PV module are the short-circuit
current Icc and the open-circuit voltage VCO. Icc and VCO change with the incident solar
irradiance f and with the ambient air temperature Ta. The short-circuit current is about
proportional to the incident solar irradiance and the open-circuit voltage increases just a
little when the solar irradiance increases. On the other hand, it is important to note that VCO
decreases with increasing module temperature which leads to a noticeable decrease in the
available maximum electrical power, in spite of a small increasing of the short-circuit
current ICC [5].
Andreev et al. [6] calculated that the photocurrent increases with the temperature at
0.1% 8CK1 due to the decreasing of the gap of the solar cell and that the open-circuit
voltage decreases at K2 mV 8CK1 between 20 and 100 8C due to a reduction of the gap
but also due to an increasing of the saturation current. These two effects lead to a decrease
for the maximum available power equal to 0.35% 8CK1. Nolay [7] estimated the same
reduction at about 0.5% 8CK1. More recently, this influence has been estimated between
K0.3 and K0.5% 8CK1 [8–11].
It is obvious that these influences on Icc and VOC have some consequences on the
electrical efficiency of the PV cell or module. The relative temperature coefficient of
crystalline silicon solar modules is in the range 0.4–0.6% 8CK1 according to Moshfegh et
Sandberg [12]. With a 13% absolute conversion efficiency this corresponds to an absolute
temperature coefficient between 0.031 and 0.046% 8CK1. Therefore, a reduction by 20%
will give an increase in efficiency between 0.6 and 1%.
Del Cueto [13] studied the performances of several photovoltaic modules using
various technologies: crystalline silicon (c-Si), Polycrystalline silicon (pc-Si), Cadmium
Telluride (CdTe), and Copper Indium Diselenide (CIS). He noted that for all c-Si, poly
c-Si and CIS modules, the changes in efficiency due to temperature dependence appear
to be in the range of absolute 1–2% over a temperature change span of 30 8C. The
dependences in % 8CK1 of the module efficiency are given in Table 1 for these
technologies [13]. We note that all these data found in the literature are in good
agreement.
For a a-Si module, the temperature coefficients of the efficiency are typically lower at
K0.1% 8CK1 (against K0.4% 8CK1 for c-Si and CIS) [14].
M. Mattei et al. / Renewable Energy 31 (2006) 553–567 555

Table 1
Effect of the temperature on the module efficiency according to Del Cueto [13]

Module c-Si c-Si c-Si pc-Si pc-Si CIS CdTe


type
% 8CK1 K0.496 K0.388 K0.427 K0.401 K0.431 K0.484 K0.035

All these effects must be considered in any model for photovoltaic module efficiency.
The most known model is given by the following equation:
h Z hr ½1 K bðTc K Tr Þ C g Log f (1)
where hr is the reference module efficiency at a PV cell temperature Tr of 25 8C and at
a solar irradiance f on the module equal to 1000 W mK2. g and b are, respectively, the
solar irradiance and temperature coefficients for the PV module. Tc is the PV cell
temperature which depends on the environmental conditions. Generally, these
parameters (Tr,hr,b,g) are given by the photovoltaic manufacturer, but g and b depend
on the material used for the PV module. Evans [15] suggested to use for the silicon bZ
0.0048 8CK1 and gZ0.12 and for a CIS module bZ0.006 8CK1. Most often this
equation is seen with gZ0 [16].
In a comparative study concerning the performances of photovoltaic/thermal solar air
collectors, Hegazy [4] used, on the basis of the study of Bergene et al. [1], Eq. (1) for the
efficiency of the PV module taking for g and b the respective values of 0 and 0.004 8CK1,
the reference efficiency being taken equal to 12.5%. This formulation (Eq. (1)) has been
also used with gZ0 in a study of a PV/thermal collector [17].
Another formulation has been used by Sandnes and Rekstad [18], the efficiency
depends linearly of the temperature:
h Z hr K mðTc K Tr Þ (2)
the temperature coefficient m has been calculated for a 41 Wp module. They found for m
equals to 0.07 and 0.1% 8CK1 for serial and parallel combinations, respectively. These
values are higher than the values found by Saidov et al. [19] with mZ0.05% 8CK1 for a
silicon module. We remark that for gZ0 Eq. (2) is identical to Eq. (1) with mZbhr.
In the above two models, Eqs. (1) and (2), the cell temperature appears, thus to be an
important parameter to study.

2.2. Experimental verification of the efficiency formulations

In this section, we want to verify Eqs. (1) and (2) which allow the calculation of the PV
module efficiency versus the PV module temperature. Although the influence of the solar
irradiance has been neglected in these two Eqs. (1) and (2), it must be kept in mind that the
influence of the solar radiation is integrated in the cell temperature because it depends
strongly on the amount of solar irradiance.
We have a photovoltaic system connected to the electrical grid with a PV array
composed of ten 85 Wp crystalline BP585F modules connected to a 700 W Sunny Boy
1000 inverter.
556 M. Mattei et al. / Renewable Energy 31 (2006) 553–567

0.14

0.135 experimental efficiency


calculated efficiency using Eq. (1)
Calculated efficiency using Eq. (2)
0.13
Electrical efficiency

0.125

0.12

0.115

0.11

0.105

0.1
1
16
31
46
61
76
91
106
121
136
151
166
181
196
211
226
241
256
271
286
301
316
331
346
361
376
391
406
421
436
451
466
Time (minutes)

Fig. 1. Comparison of the two models of efficiency with experimental data.

We collected each minute five data: PV array voltage and current, solar irradiance,
ambient and module temperature. We calculated the experimental efficiency and compared
it with efficiency estimated using Eqs. (1) and (2) taking as cell temperature the temperature
measured at the backside of the PV module as it is often realized in such a case [11,13].
We took as values for the coefficients hrZ0.125, TrZ25 8C, bZ0.0044 8CK1 and mZ
0.05% 8CK1. Obviously, if we have taken mZbhr, the two curves should be identical.
The experimental efficiency and the two modelled ones are presented in Fig. 1 for a
particular day (2001, April 21). The model corresponding to Eq. (1) gives better results.
The other model does not follow the variation of the efficiency.
From Eq. (1), we calculate the electrical power produced by the PV array and we
present in Fig. 2 the result of this model for a particular day (2001, April 29).
We note that the model is a good accordance with the experimental data. We calculated,
on the basis of 1 month, the mean bias error (MBE), the root mean square error (RMSE) and
the correlation coefficient (CC) for the efficiency and electrical power model (Table 2).
We consider that the performances of the model are satisfying considering the
simplicity of the model and the low number of parameters used.

3. Cell temperature

3.1. Some models of cell temperatures

This temperature influences the I–V characteristics and consequently the electrical
efficiency of the PV module.
M. Mattei et al. / Renewable Energy 31 (2006) 553–567 557

700

600
PV Array Electrical Power (W)

500

400
Experimental electrical power produced by the PV array

Modelled electrical power produced by the PV array


300

200

100

0
1 41 81 121 161 201 241 281 321 361 401 441 481 521 561 601 641 681 721
Time (minutes)

Fig. 2. Experimental and modelled PV array electrical power.

The most common manner to determine the cell temperature Tc consists in using the
Normal Operating Cell Temperature (NOCT) [7]. The value of this parameter is given by
the PV module manufacturer. Tc is then dependent on the ambient temperature Ta and on
the solar irradiance f according to Eq. (3):
f
Tc Z Ta C ðNOCT K 20 + CÞ (3)
800
This simple method yields satisfying results if the PV modules are not roof-integrated.
A complete definition of NOCT and the conditions of determination of this parameter are
presented in Refs. [20,21]. NOCT is calculated for a wind speed at a PV module height of
nZ1 m sK1, an ambient temperature TaZ20 8C and a hemispherical irradiance fZ
800 W mK2. NOCT depends strongly of the type of encapsulation of the PV module.
An energy balance on a PV module has been realized with some hypothesis:

† we neglect the difference of temperatures between photovoltaic cells and the cover
(glass);
† we consider that the temperature is uniform in the panel;
† the radiative exchanges are considered as negligible.

Table 2
Values of statistical parameters for efficiency and electrical power models

MBE RMSE CC
Efficiency K0.0160 0.033 0.856
Electrical power K0.0132 W 35 W 0.882
558 M. Mattei et al. / Renewable Energy 31 (2006) 553–567

For a solar irradiance f, the part crossing the glass is tf where t is the transmittance of
the cover system for beam and diffuse radiation and the part absorbed by the photovoltaic
cells is atf with a the absorption coefficient of the cells. The energy losses are:

† the losses due to the electrical power produced by the photovoltaic cells: hf
† the thermal losses from the collector to the surroundings: UPV(TcKTa).

The energy balance is:


atf Z hf C UPV ðTc K Ta Þ (4)
such a model has been used by Furler et al. [22] and Sandnes and Rekstad [18].
If we add Eq. (1) with gZ0 in Eq. (4), we obtain:
UPV Ta C f½ðatÞ K hr K bhr Tr 
Tc Z (5)
UPV K bhr f
or if we introduce Eq. (2) in Eq. (4), we obtain:
UPV Ta C f½ðatÞ K hr K mTr 
Tc Z (6)
UPV K mf
According to Furler et al. [22], (at)/UPVZ0.0325 8C m2 WK1 and according to
Sandnes et Rekstad [18] (at)Z0.9 and UPVZ28.8 W 8CK1 mK2, the two values of UPV
are in a good agreement. However, the prediction of UPV can be improved by taking into
account the effect of wind speed which is a meteorological variable influencing this loss
coefficient UPV. Several correlations are available to compute UPV versus the wind speed v.
A study performed by Jones and Underwood [23] reveals a considerable range of
values for the forced convection coefficient; for a wind speed of 1 m sK1 the value of hc is
1.2 W mK2 8CK1 [24], 5.8 W mK2 8CK1 [25], 9.1 W m2 8CK1 [26] or 9.6 W mK2 8CK1
[27].
Because of the wide discrepancies in the value for hc, it is difficult to choose a particular
value. Duffie and Beckman [28] suggest for hc to use the expression given by McAdams
[29] for flat plates exposed to outside winds:
hc Z 5:67 C 3:86v (7)
Nolay [7] used the following relation:
hc Z 5:82 C 4:07v (8)
According to Cole and Sturrock [30], the convective surface heat transfer coefficient is
strongly dependent upon the wind direction and whether the subject surface is on
windward or leeward side and can be expressed by:
hc Z 11:4 C 5:7v (9)
for a windward surface
hc Z 5:7 (10)
for a leeward surface
M. Mattei et al. / Renewable Energy 31 (2006) 553–567 559

UPV is the heat exchange coefficient corresponding to the total surface area of the
module, i.e. two times the surface area corresponding to hc because the heat is lost by the
two faces of the photovoltaic module (lateral surfaces are neglected), thus,
UPV Z 11:34 C 7:72v (11)
if Eq. (7) is used;
UPV Z 11:64 C 8:14v (12)
if Eq. (8) is used;
UPV Z 17:1 C 5:7v (13)
if Eqs. (9) and (10) are used.
For a wind speed equal to 1 m/s, the respective values of UPV are UPVZ19.06 mK2 8CK1
from Eq. (11), 19.78 W mK2 8CK1 from Eq. (12), 22.8 W mK2 8CK1 from Eq. (13) and
UpvZconstantZ28.8 W mK2 8CK1 [18,22] without taking into account the wind speed.
Barker and Norton [31] proposed a formulation for the energy balance established by
Ingersoll [32] and taking all the heat transfers including radiative one and differentiating
the heat transfer coefficients according to the exchange surface. This model uses an
important number of coefficients and a correlation table is needed to determine some
parameters.

3.2. Presentation of the tested photovoltaic module

The photovoltaic module tested is a Photowatt PWX 500 using multi-crystalline


technology with a thickness of 0.2 mm. The encapsulation of cells is made between two
sheets of tempered glass with high transmittance. The dimension of the module is
1042 mm!462 mm!39 mm. The peak power at a junction temperature equal to 25 8C is
49 W at G10%.
Eight thermal sensors have been integrated into the module during its manufacturing:
the first one measures the temperature on the back surface of the glass cover (no. 1), the
second one measures the temperature of the back surface of the module (no. 8) and the
other six sensors measure the temperature on six points on the back surfaces of cells (no.
2–no. 7). The positions of the temperature sensors are presented in Fig. 3.

3.3. Evolution of the different temperatures

A temperature sensor was faulty, just five temperatures on the back surfaces of the PV
cells have been measured. We wanted to observe the difference of temperature between
the five points of measurements on the photovoltaic module, thus we plotted in Fig. 4 the
evolution of these five temperatures and the average temperature (calculated from these
values) for three consecutive days.
We note that the temperature gradient in the PV module is very low. This fact confirms
the second assumption used in the energy balance, i.e. the temperature is uniform in the
photovoltaic panel. Thus, we can consider an average temperature of the photovoltaic
560 M. Mattei et al. / Renewable Energy 31 (2006) 553–567

Glass cover
Temperature sensor
n°1
Photovoltaic cells
Temperature sensors
n°2 to n°7
Back glass
Temperature sensor
n°8
Temperature sensor

Fig. 3. Presentation of the photovoltaic module and position of the temperature sensors.

module and it is this temperature which has been used to validate the photovoltaic cell
temperature.

3.4. Estimation of the module temperature for a wind speed of 1 m/s

From the experimental values measured for a wind speed equal to 1 m/s, we calculated
the optimal values for (at) and UPV. We found (at)Z0.81 and UPVZ28.9 W mK2 8C. In
the literature, we found for (at), 0.855 for Photowatt [33], 0.81 for Hegazy [34], 0.875 for

Fig. 4. Evolution of the cells temperature in the PV module.


M. Mattei et al. / Renewable Energy 31 (2006) 553–567 561

700 45
Estimated temperature using NOCT method
Estimated temperature using energy balance and a constant Upv
Estimated temperature using energy balance and Upv calculated by Eq. (11)
Estimated temperature using energy balance and Upv calculated by Eq. (12) 40
600 Estimated temperature using energy balance and Upv calculated by Eq. (13)
Estimated temperature using optimal calculated values
Solar irradiance 35
Ambient temperature
500 Measured PV module temperature
Solar irradiance (W/m2)

30

Temperatures (°C)
400 25

300 20

15
200
10

100
5

0 0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26
Acquisition number

Fig. 5. Verification of the PV module temperature models for a wind speed of 1 m/s.

Brogren et al. [35] and 0.9 for Sandnes and Rekstad [18]. Thus, we see that your optimal
value is in accordance with the literature.
We estimated the PV module temperature using the models presented in Section 3.1:

† NOCT model (Eq. (3))


† Model based on the energy balance using respectively for UPV the Eq. (11) [28], Eq.
(12) [7] and Eq. (13) [30],
† Model based on energy balance and your optimized values of (at) et UPV value

and we compared these estimated values with the average module temperature
calculated from measured temperatures. The comparison between experimental and
modelled temperature for a wind speed equal to 1 m/s is illustrated in Fig. 5.
The values of the statistical coefficients calculated for this comparison between
experimental data and the five models are given in Table 3.
It appears that the best results are obtained for the model based on the energy balance and
your optimal coefficient but the second is the model using a constant value for UPV. The model
taking into account the wind speed and being the more appropriate is the model using for UPV
the correlation of Cole and Sturrock [30]. The classical formulation using the NOCT leads to
satisfying results. We want to test these models whatever the value of the wind speed is.

3.5. Validation of the models for variable wind speeds

We calculated, from experimental data and for all wind speeds, the optimal values for
the thermal coefficient UPV considering for the optimal value of (at) the previous
calculated value equal to 0.81.
562 M. Mattei et al. / Renewable Energy 31 (2006) 553–567

Table 3
Values of the statistical coefficients for various models (wind speed of 1 m/s)

NOCT Energy balance


Constant UPV UPV UPV Optimised
UPV Eq. (11) Eq. (12) Eq. (13) UPV and at
MBE (8C) 2.41 1.37 4.06 3.72 2.99 0.61
RMSE (8C) 3.03 2.19 4.78 4.38 3.57 1.96
MBE/average value 0.137 0.078 0.230 0.211 0.169 0.035
of temperature
RMSE/average value 0.172 0.124 0.271 0.249 0.202 0.111
of temperature
CC 0.949 0.964 0.938 0.942 0.948 0.973

Two expressions have been found for UPVZaCbv:

† the first one considering that the expression of UPV have to verify the previous optimal
value found for vZ1 m sK1, i.e. UPVZ28.9 W mK2 8CK1. This constraint implies to
have aCbZ28.9.
† the second expression without this constraint.

The first expression is:

UPV Z 26:6 C 2:3v (14)

with for vZ1 m/s UPVZ28.9 W mK2 8CK1


the second expression is:

UPV Z 24:1 C 2:9v (15)

with for vZ1 m/s UPVZ27 W mK2 8CK1.


During several consecutive days, we observed the evolution of the PV module
temperature and we estimated it using the models previously described. We plotted, in
Fig. 6., as an example for 2 days (12/02 and 12/03), the evolution of the ambient
temperature, of the solar irradiance, of the wind speed at the PV module level and of the
module temperature. On the same figure, the modelled module temperatures are also
presented.
The module temperature estimated respectively by the NOCT model and by the
energy balance usOg a constant thermal coefficient do not follow the variation of the
experimental PV module temperature, the influence of the wind is consequently not
negligible.
The validation of these models have been established on the basis of 10 days of data
collected each minute and the values of the statistical coefficients are presented in
Table 4.
The NOCT model is not in accordance with experimental data because this model has
been conceived only for particular environmental conditions. The models based on the
energy balance and taking into account the variation of the wind speed lead to better
performances excepted for the model using the Duffie and Beckman correlation [28].
M. Mattei et al. / Renewable Energy 31 (2006) 553–567 563

Fig. 6. Evolution of the various meteorological parameters and comparison between modelled and experimental
PV module temperature.

It is obvious that the simulation using our optimized coefficients for (at) and UPV,
(respectively Eq. (14) and Eq. (15)) gives the best result. The performance of the model
using the energy balance with the heat transfer coefficient calculated by the Cole and
Sturrock correlation [30] is in good accordance with the experimental data as previously
noted for a constant wind speed equal to 1 m sK1.
We must keep in mind that the energy balance for the PV module used to determine
Eq. (5) did not take into account radiative transfer but only convective one (natural and
forced). We note that such a simple model gives satisfying results.

Table 4
Values of the statistical coefficients for various models (variable wind speed)

NOCT Energy balance


Constant UPV UPV UPV UPV UPV
UPV Eq. (11) Eq. (12) Eq. (13) Eq. (14) Eq. (15)
MBE (8C) 2.90 2.28 2.10 1.95 1.99 1.39 1.49
RMSE (8C) 3.93 3.03 3.47 3.32 2.76 2.26 2.24
MBE/average value 0.256 0.202 0.186 0.173 0.176 0.123 0.131
of temperature
RMSE/average 0.347 0.267 0.306 0.293 0.244 0.200 0.198
value of temperature
CC 0.973 0.976 0.954 0.954 0.976 0.986 0.986
564 M. Mattei et al. / Renewable Energy 31 (2006) 553–567

4. Combination of the two modellings

We combined the best simplified modelling of the photovoltaic module temperature


with the best modelling of the electrical efficiency. Thus, we introduced Eq. (5) for Tc and
Eq. (15) for UPV into Eq. (1) and we computed the PV array output power which is
compared with the experimental electrical power.
We present in Fig. 7 a comparison, during 1 day, between the modelled and
experimental PV array output power and we plotted the essential meteorological
parameters.
The agreement between the modelled and experimental electrical power is good and
allows to study the influence of meteorological parameters on the electrical power
produced by a photovoltaic panel.
As reported in Section 2, the PV array output power decreases with the temperature.
This decreasing was estimated at K0.35% 8CK1 by Andreev et al. [6], K0.5% 8CK1 by
Nolay [7], between K0.3 and K0.5% 8CK1 by Krauter et al. [11], between K0.4 and
K0.6% 8C K1 for Moshfegh and Sandberg [12] and between K0.388% and
K0.496% 8CK1 for m-Si PV modules and between K0.401 and K0.431% 8CK1 for
p-Si modules according to Del Cueto [13].
We found with our model a decrease between 0.401% 8CK1 (vZ10 m sK1, fZ
250 W mK2 and TaZ1 8C) and 0.563% 8CK1 (vZ1 m/s, fZ650 W mK2 and TaZ50 8C).
These results are in accordance with the literature.
The influence of the wind speed takes place in the heat transfer coefficient value. We
plotted the PV array output power for a given ambient temperature and a given solar

20 1200
Solar irradiance
ambient temperature
Wind speed
Solar irradiance (W/m2) - PV output power (W)
18
PV array ouput power (Estimated)
PV array output power (Experimental) 1000
Temperature (°C) - Wind speed (m/s)

16

14
800
12

10 600

8
400
6

4
200
2

0 0
6:55 7:40 8:25 9:10 9:55 10:40 11:25 12:10 12:55 13:40 14:25 15:10 15:55 16:40 17:25 18:10 18:55 19:40 20:25
Time

Fig. 7. Verification of the PV array output power model for one particular day.
M. Mattei et al. / Renewable Energy 31 (2006) 553–567 565

80
Electrical PV array output power per m2 (W/m2)

70

T= 20 ˚C, E= 250W/m² T=40˚C, E=250W/m²


60 T= 20 ˚C, E= 650W/m² T=40˚C, E=650W/m²

50

40

30

20
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25

Wind speed (m/s)

Fig. 8. Influence of the wind speed on the PV module production.

irradiance versus the wind speed in Fig. 8. This influence is felt particularly for low wind
speeds, because after about 10 m/s the impact of an increase of the wind speed seems to be
less important.

5. Conclusion

Some models of module temperature are proposed, the first one using the notion of
NOCT and the other ones using an energy balance on the PV module taking into account
the heat convective coefficient. The five last models differ by the value of the heat transfer
coefficient taken constant or estimated versus the wind speed by three different
correlations found in the literature.
The estimated module temperatures have been compared with experimental data of a
PV module specially equipped of thermal sensors. Statistical parameters have been
computed to determine the validity of each model.
The best model is based on the energy balance and uses for the heat transfer a relation
calculated from experimental data, the coefficient (at) has been also optimised with a
value of 0.81. With this optimised model the root mean square error is 2.24 8C, i.e. a
relative RMSE of 19.8%. The same methodology using the correlation of Cole and
Sturrock [30] gave satisfying results (RMSEZ2.76 8C). This error is not very low but we
must keep in mind that this model is a very simplified model.
In the same way, we tested two models of PV modules electrical efficiency taking into
account the module temperature. A grid connected photovoltaic system installed in our
laboratory allows us to compare these models with experimental results. We found good
566 M. Mattei et al. / Renewable Energy 31 (2006) 553–567

performances for the first model and we decided to couple it with the model of module
temperature. Thus, we determined the PV array output power for various ambient
temperatures, solar irradiances and wind speeds. The influence of these environmental
parameters have been compared with the literature and the results obtained by the model
are in accordance with previous works.

References

[1] Bergene T, Lovik O. Model calculations on a flat-plate solar heat collector with integrated solar cells. Solar
Energy 1995;55(6):453–62.
[2] Huang BJ, Lin TH, Hung WC, Sun FS. Performance evaluation of solar photovoltaic/thermal systems. Solar
Energy 2001;70(5):443–8.
[3] Tripanagnostopoulos Y, Nousia TH, Souliotis M, Yianoulis P. Hybrid photovoltaic/thermal solar systems.
Solar Energy 2002;72(3):217–34.
[4] Hegazy AA. Comparative study of the performances of four photovoltaic/thermal solar air collectors.
Energy Conversion Manage 2000;41:861–81.
[5] Buchet E. Etude du dimensionnement et développement d’un logiciel d’aide à la conception de systèmes de
production d’énergie utilisant la conversion photovoltaı̈que de l’énergie solaire. MS Thesis, University of
Saint Jérôme, Marseille, France; 1988.
[6] Andreev VM, Grilikhes VA, Rumyantsev VD. Photovoltaic conversion of concentrated sunlight. London:
Wiley; 1967 [ISBN: 0471967653].
[7] Nolay P. Développement d’une méthode générale d’analyse des systèmes photovoltaı̈ques. MS Thesis,
Ecole des Mines, Sophia-Antipolis, France; 1987.
[8] Emery K, Burdick J, Caiyem Y, Dunlavy D, Field H, Kroposki B, et al. Temperature dependance of
photovoltaic cells, modules and systems. In: Proceeding of the 25th IEEE PV specialists conference,
Washington, USA; 13–19 May 1996, p. 1275–8.
[9] King DL, Eckert PE. Characterizing (rating) the performance of large photovoltaic arrays for all operating
conditions. In: Proceeding of the 25th IEEE PV specialists conference, Washington, USA; 13–19 May 1996,
p. 1385–8.
[10] Wilshaw AR, Bates JR, Pearsall NM. Photovoltaic module operating temperature effects. In: Proceeding of
Eurosun’96, Munich, Germany; 1996, p. 940–4.
[11] Krauter S, Araujo RG, Schroer S, Hanitsch R, Salhi MJ, Triebel C, Lemoine R. Combined photovoltaic and
solar thermal systems for facade integration and building insulation. Solar Energy 1999;67(4–6):239–48.
[12] Moshfegh B, Sandberg M. Flow and heat transfer in the air gap behind photovoltaic panels. Renewable
Sustainable Energy Rev 1998;2:287–301.
[13] Del Cueto JA. Comparison of energy production and performance from flat-plate photovoltaic module
technologies deployed at fixed tilt. In: Proceeding of the 29th IEEE PV specialists conference, New Orleans,
USA; 20–24 May 2002.
[14] Photon International. Market survey on solar modules; February 2004, p. 46–55.
[15] Evans DL. Simplified method for predicting photovoltaic array output. Solar Energy 1981;27(6):555–60.
[16] Evans DL, Florschuetz LW. Cost studies on terrestrial photovoltaic power systems with sunlight
concentration. Solar Energy 1977;19:255.
[17] Garg HP, Adhikari RS. Conventional hybrid photovoltaic/thermal (PV/T) air heating collector: steady-sate
simulation. Renewable Energy 1997;11(3):363–85.
[18] Sandnes B, Rekstad J. A photovoltaic/thermal (PV/T) collector with a polymer absorber plate, experimental
study and analytical model. Solar Energy 2002;72(1):63–73.
[19] Saidov MS, Abdul’nabi ZM, Bilyalov RR, Saidov AS. Temperature characteristics of silicon solar cells.
Appl Solar Energy 1995;31(6):84–8.
[20] ASTM. Standard test methods for electrical performance of nonconcentrator terrestrial photovoltaic
modules ad arrays using reference cells, standard E1036. West Conshohocken, PA: American Society for
Testing and Materials; 1998.
M. Mattei et al. / Renewable Energy 31 (2006) 553–567 567

[21] Myers DR, Emery K, Gueymard C. Revising and validating spectral irradiance reference standards for
photovoltaic performance. In: ASES/ASME solar 2002 conference proceeding, Reno, Nevada; 15–20 June
2002.
[22] Furler G, Beckman WA, Klein SA. Modeling of a photovoltaic powered refrigeration system. Personal
communication; 1994.
[23] Jones AD, Underwood CP. A thermal model for photovoltaic systems. Solar Energy 2001;70(4):349–59.
[24] ASHRAE. ASHRAE handbook: fundamentals 1989.
[25] Anis WR, Mertens RP, Van Overstraeten R. Calculation of solar cell operating temperature in a flat plate PV
array. In: Proceeding of the 5th PV solar energy conference; 1983, p. 520–4.
[26] Schott T. Operational temperatures of PV modules. In: Proceeding of the 6th PV solar energy conference;
1985, p. 392–6.
[27] Pratt AW. Heat transmission in buildings. London: Wiley; 1981.
[28] Duffie JA, Beckman WA. Solar energy thermal processes. London: Wiley; 1974 [ISBN 0-471-22371-9].
[29] McAdams WC. Heat transmission. 3rd ed. New York: McGraw-Hill; 1954.
[30] Cole RJ, Sturrock NS. The convective heat exchange at the external surface of buildings. Building Environ
1977;12:207–14.
[31] Barker G, Norton P. Predicting long-term performance of photovoltaic arrays using short-term test data and
an annual simulation tool. NREL Report, CP-550-33601 presented at the Solar 2003 Conference: America’s
Secure Energy, Austin, Texas; 2003.
[32] Ingersoll JG. Simplified calculation of solar cell temperatures in terrestrial photovoltaic arrays. J Solar Eng
1986;108:95–101.
[33] Photowatt SA. Personnal communication; 2000.
[34] Hegazy AA. Comparative study of the performances of four photovoltaic/thermal solar air collectors.
Energy Conversion Manage 2000;41:861–81.
[35] Brogren M, Nostell P, Karlsson B. Optical efficiency of a PV-thermal hybrid CPC module for high latitudes.
Solar Energy 2000;69(Suppl):173–85.

You might also like