You are on page 1of 11

Solar Energy 269 (2024) 112345

Contents lists available at ScienceDirect

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

Reducing the operating temperature and improving the efficiency of PV


modules using guiding vanes
Raed I. Bourisli a,b ,∗, Fatema A. Zaher b,c , Bader S. Aldalali d
a Mechanical Engineering Department, Seattle University, 901 12th Ave, Seattle, WA 98122, USA
b
Mechanical Engineering Department, Kuwait University, PO Box 5969, Safat 13060, Kuwait
c
Kuwait Packing Materials Manufacturing Company, 82nd Street, Subhan 13126, Kuwait
d
Electrical Engineering Department, Kuwait University, PO Box 5969, Safat 13060, Kuwait

ARTICLE INFO ABSTRACT

Keywords: Reduction of the operating temperature of photovoltaic cells can significantly increase their efficiency. An
PV module effective way of reducing this temperature is to redirect ambient air to carry some of the thermal energy away
Passive cooling from the modules. In this work, guiding vanes placed under the modules are used to accomplish that. The
Guiding vanes
shapes, sizes, and locations of the proposed vanes are optimized to give the greatest temperature reduction
Efficiency
based on realistic conditions found in the Kabd Desert of Kuwait. A 2D finite element model, validated using
Temperature reduction
an experimental PV module built in the Kabd desert of Kuwait. Optimum vanes were found for multiple rows
of modules, resulting in temperature reductions of 8–12.6 ◦ C for the representative late-spring day tested. With
such decreases in operating temperature, a typical solar module operating in such conditions can increase in
efficiency roughly from 17% to 18%. Specific airflow patterns around the optimized vanes/modules were found
to be a key factor. The vanes gradually divert increasing fractions of wind towards the backs of the modules.
The fractions of air flowing above the vanes, onto the back of the module, compared to those going below the
vane, continuing on the next module, go from 28%–72% to 43%–57% to 64%–36% to finally 73%–27%. The
vanes optimally apportion the cooling potential, thereby lowering the maximum temperature across the entire
array. The results demonstrate and quantify the efficacy of such guiding vanes in lowering the temperature
and increasing the efficiency of PV modules. It also extends the area of passive cooling of PV modules and
points towards possible future directions.

1. Introduction and background increase of 1 ◦ C can decrease the efficiency by more than 0.3% [2–4].
Many factors affect module temperature and thereby efficiency,
Photovoltaic (PV) cells convert incident solar irradiation to electri- such as dust, shading, condensation, installation height, wind speed,
cal energy. A PV cell is made of semiconductor materials that absorb and wind direction, among others. The rise in the operating temper-
the photons emitted by the sun and generate a flow of electrons, or, ature of the modules can also have other detrimental effects, besides
electric current. The absolute amount of electric current supplied by a reducing the energy output, such as module degradation and reduced
PV cell is naturally determined by the amount of irradiation it receives. lifetime [3,5]. Having an effective cooling system for the PV modules
Its efficiency is the ratio of the amount of irradiation it receives to the is therefore a major key to enhancing their performance [6].
electrical power it produces. Absolute amounts are governed by several Cooling techniques can either be active or passive, and both could
factors. use some type of fluid to move the collected energy away from the
PV cells typically have low efficiencies, often converting only 14%– modules. Passive cooling techniques that use air are the most prevalent
25% of the solar irradiation they receive to useful electric power [1]. because of their reliability and cost-effectiveness. They also do not
The remainder of the incident radiative energy shows up as heat–in the involve external power sources, mechanical devices, moving parts, or
cells, in the module structure and in the electric wiring. This heat can plumbing. The goal of all these cooling techniques is to enhance the
substantially raise the temperature of the PV modules, reducing their performance of PV modules. Hasanuzzaman et al. [2] and Sato and
efficiency and affecting their performance. Studies have shown that an Yamada [7] provide a comprehensive overview of common cooling

∗ Corresponding author at: Mechanical Engineering Department, Kuwait University, PO Box 5969, Safat 13060, Kuwait.
E-mail addresses: bourislr@seattleu.edu (R.I. Bourisli), fatima@kuwaitpack.com (F.A. Zaher), bader.aldalali@ku.edu.kw (B.S. Aldalali).

https://doi.org/10.1016/j.solener.2024.112345
Received 13 October 2023; Received in revised form 15 December 2023; Accepted 12 January 2024
Available online 28 January 2024
0038-092X/© 2024 International Solar Energy Society. Published by Elsevier Ltd. All rights reserved.
R.I. Bourisli et al. Solar Energy 269 (2024) 112345

techniques in use nowadays and their general quantitative perfor- in the outdoor environment and is usually not significant to warrant
mance attributes and limitations. Some of these techniques are briefly independent analysis, unless for the sake of comparison [24]. Most
reviewed below. analyses are done for forced convection—turbulent in particular [15].
Using water, via evaporative cooling [8–10] and phase-change ma- Since the top surface of a (monofacial) PV module is exposed to
terials (PCM) [11–13] are among the best cooling techniques available, the wind, the rise in temperature and the associated heat spots and
in terms of the temperature reduction achieved. However, these tech- stagnation regions usually occur behind the modules. A few exceptions
niques are rarely applied on a mass-scale due to the accompanying had compared free, laminar and turbulent cases [18]. The author
need for (pumping) power sources, relatively intricate plumbing, good optimized the geometry of the underlying channel and found it to
maintenance programs, and initial capital. To avoid these complica- be independent of slight changes in module inclination angle. More
tions, many researchers have turned to using atmospheric air to cool the recently, Zhang et al. [25] presented a porous channel to cool PV
modules. Skoplaki and Palyvos [14] surveyed a few dozen correlations modules. They conducted optimization of a number of parameters, with
for cell/module temperature based on radiative flux, efficiency, glass and without a fan, pertaining to the porous structure and the channel.
properties, and many other environmental parameters. They concluded These included flow direction, number of holes, sizes and distribution,
that no general guidelines can be established regarding the operating and other parameters, at a number of operating conditions. An increase
temperature. As a mere example, Fouladi et al. [15] investigated the of 2.7% in power conversion efficiency was reported.
dependence on wind turbulence intensity. Even though the wind in the In the current work, a set of optimized guiding vanes is used to
guide ambient wind to flow over the back surfaces of PV modules,
experiments was controlled, they stated that the amount of cooling was
reducing their temperatures and enhancing their efficiency. It is the
related to a number of other factors, such as the design and material
backs of modules where air typically has low velocities, resulting in
of the system, in addition to wind velocity and turbulence intensity.
heat build ups, higher temperatures, and, ultimately, low efficiencies.
This confirms the intricacy involved in assessing the wind potential in
The vanes therefore are placed towards the lower end of the modules,
cooling PV modules.
to collect as much incoming air as possible, as efficiently as possible.
In general, PV cooling techniques that use air can be classified as
Vanes of different sizes and shapes are placed below the module.
either passive or active. They can range from simply allowing natural
Then, an array of modules is modeled, and the guiding vanes are
convection to move the air around the module, to using mechanical
consecutively affixed to all of them. The locations, sizes, and radii of
power to circulate the air, removing thermal energy from the module
curvature of these vanes are the subject of a sequential optimization
and its vicinity. Teo et al. [16] discussed the difference between the study, with the objective of minimizing the temperature of the modules
passive (natural convection only) and active cases, using different to increase their efficiency. The next section describes the problem
blowers and settings. A 45% improvement in efficiency (from 8.6% mathematically and dynamically, including the various assumptions.
to 12.5%) was reported. In one of the earliest works on the sub- The validation of the numerical model is discussed in Section 3. Results
ject, Brinkworth et al. [17], Brinkworth [18] showed that the efficiency are presented and discussed in Section 4, while conclusions are drawn
of PV modules could be cut by one third if the temperature was left to in Section 5.
rise, compared to Standard Test Conditions (STC). Cooling by air has
seen steadily increasing attention, however, over the past two decades 2. Problem description
due to innovative management of the fluid flow systems [19].
Another common technique of enhancing heat transfer from mod- 2.1. PV system design
ules is simply using fins, i.e., extending the heat transfer area. This
has been done using fins of multiple types (high efficiency fins, high A 2D cross-sectional model of the PV glass-glass mini-module con-
conductivity fins, etc.), shapes (straight fins, pin fins, etc.), and sizes. El- sisting of 9 cells (3 × 3) is created, based on the physical 0.16 × 0.16
breki et al. [20] installed finned heat sinks on the back of the modules m2 cells. The thickness of the PV module is 7 mm. An adjusted set
and looked at the reduction of the module temperature using planar of glass properties was used for the PV mini-module that incorporates
reflectors coupled with longitudinal lapping fins. Compared to the from the top: a 3.2-mm glass, a 0.45-mm clear encapsulant, a 0.18-mm
bare PV module, significant reductions in operating temperature were crystalline silicon based n-type passivated emitter rear totally diffused
observed. The number of fins was also primitively optimized in the (NPERT) solar cell, a 0.45-mm white encapsulant, and a 3.2-mm glass
process. Some life cycle cost analysis was done as well, and it showed bottom. The PV module sits on a frame elevated 0.25 m above ground,
payback periods halved, compared to bare PV modules. Such tech- due south, with a 30-degree tilt. Figs. 1(a) and 1(b) show the 3D and 2D
niques have recently been shown to be very effective in bifacial PV rendering of the module, respectively, while Fig. 2 shows a schematic of
modules [21]. the problem with key geometrical and boundary conditions. In that last
Many researchers combined PV modules with other devices that figure, 𝐿1 = 1.5 m is the upstream section length that guarantees fully
provide airflow, ventilation, auxiliary power, and/or other functions, turbulent boundary layer over the entire module. Meanwhile, 𝐿2 = 15
but are primarily used for the purpose of cooling the module. Jet m is the downstream section length that results in limited downstream
impingement, for example, was used to focus airflow onto the PV eddy disturbance at the domain outlet. The ground clearance, 𝐻1 = 0.25
m, mimics the actual module height while 𝐻2 = 8 m is to guarantee
modules and increase its momentum, and thus cooling capacity. Royne
uniform flow at the top.
and Dey [22] studied the optimum number and locations of nozzles,
and the gain realized with the associated temperature reduction of the
2.2. Governing equations
module, all while considering the Nusselt number over the module
surfaces and the pumping power required. In light of the experimen-
The fluid and energy flows are governed by the conservation of
tal and numerical results obtained, they concluded that optimization mass, momentum and energy equations. The continuity equation rep-
can be done based on the average heat transfer coefficient, with less resents the conservation of mass in differential form. At the typical
emphasis on temperature uniformity. Valeh-e-Sheydaand et al. [23] temperatures encountered in PV applications, density (𝜌) is not a strong
coupled the cooling function with a wind-driven ventilator. The total function of temperature [26]. The flow can therefore be approximated
generated power increased by around 46.5%, with the PV module as incompressible, or, 𝜕𝜌∕𝜕𝑇 ≅ 0, so the change in density with
accounting for more than 90% of that. The remainder came from the temperature is assumed negligible. The steady, incompressible, two-
ventilator-connected dynamo. dimensional of the continuity equation in Cartesian coordinates reduces
The most prevalent way of cooling PV modules, however, remains to,
to be using atmospheric air, be it via free, forced or mixed (free + 𝜕𝑢 𝜕𝑣
forced) convection. Pure free (natural) convection does not often exist + =0 (1)
𝜕𝑥 𝜕𝑦

2
R.I. Bourisli et al. Solar Energy 269 (2024) 112345

𝜕𝑢
where 𝑃 = 𝜏𝑖𝑗 𝜕𝑥𝑖 is the production term. The following revised model
𝑗
constants were used,
13 1 1
𝛼= , 𝜎𝑘 = , 𝜎𝑑 = , 𝛽 = 𝛽0 𝑓𝛽 ,
25 2 2
13 9 1 + 70𝜒𝜔
𝛽 ∗ = 𝛽0∗ 𝑓𝛽 , 𝛽0 = , 𝛽0∗ = , 𝑓𝛽 = ,
125 100 1 + 80𝜒𝜔
|𝛺 𝛺 𝛺 | ⎧
| 𝑖𝑗 𝑗𝑘 𝑘𝑖 | ⎪ 1 𝜒𝑘 ≤ 0
𝜒𝜔 = || ( ) |,
| 𝑓𝛽 ∗ = ⎨ 1+680𝜒𝑘2 ,
| 𝛽∗𝜔 3 | ⎪ 𝜒𝑘 > 0
| 0 | ⎩
1+400𝜒𝑘2

|𝛺 𝛺 𝛺 |
| 𝑖𝑗 𝑗𝑘 𝑘𝑖 |
𝜒𝑘 = || ( ) |
|
| 𝛽∗𝜔 3 |
| 0 |
where 𝛺𝑖𝑗 and 𝑆𝑖𝑗 and the mean rotation-rate and the mean strain-rate
tensors, respectively.
Neglecting the kinetic and potential energies and viscous dissipation
in the fluid, the conservation of energy in the fluid simplifies to,
( ) ( 2 )
𝜕𝑇 𝜕𝑇 𝜕 𝑇 𝜕2 𝑇
𝜌𝑐𝑝 𝑢 +𝑣 = 𝑘𝑓 (6)
𝜕𝑥 𝜕𝑦 𝜕𝑥2 𝜕𝑦2
where 𝑐𝑝 is the specific heat at constant pressure, and 𝑘𝑓 is the thermal
conductivity of the fluid (air). Conversely, in the solid PV cell (glass,
silicon, encapsulant, etc.), the energy is governed by,
𝜕2 𝑇 𝜕2 𝑇
+ =0 (7)
𝜕𝑥2 𝜕𝑦2
The finite element solutions of the fluid and energy flow problem were
obtained using [29].

Fig. 1. The simulated PV module. All dimensions are in meters.


2.3. Flow assumptions and boundary conditions

The flow is assumed to be steady and incompressible, with constant


transport and thermodynamic properties. A constant velocity boundary
condition is prescribed for the left domain boundary with medium
(0.5) turbulent intensity and geometry-based length scale. The top
and right boundaries are open boundaries with zero normal stress. All
other surfaces (ground, modules and vanes) are no-slip surfaces. The
aforementioned 2-dimensional approximation is used since the change
in wind direction with respect to the azimuth is neglected. Resistance to
conduction within the module was found to be negligible, making the
temperatures of the top and bottom surfaces of the module effectively
Fig. 2. The domain, PV module and basic boundary conditions. the same.
For the temperature boundary condition at the surface of the mod-
ules, temperatures at 4 different locations around the back surface of
The 𝑥- and 𝑦-components of the incompressible Navier–Strokes the 9-cell-module were noted—Figs. 3(a) and 3(b). The global incident
equations, describing the conservation of momentum, are, plane of array (GIPOA) irradiance was recorded by the nearby pyra-
( ) ( 2 ) nometer at the testing rig at Kabd over many months—Fig. 3(c). The
𝜕𝑢 𝜕𝑣 𝜕𝑝 𝜕 𝑢 𝜕2 𝑢
𝜌 + =− +𝜇 + (2) data logging system records current–voltage (I–V) data every 10 min,
𝜕𝑥 𝜕𝑦 𝜕𝑥 𝜕𝑥2 𝜕𝑦2 while irradiance, ambient and module temperature are recorded with a
( ) ( 2 )
𝜕𝑣 𝜕𝑣 𝜕𝑝 𝜕 𝑣 𝜕2 𝑣 30-second resolution. A specific instant of time was chosen, represent-
𝜌 + =− +𝜇 + (3)
𝜕𝑥 𝜕𝑦 𝜕𝑦 𝜕𝑥2 𝜕𝑦2 ing the 90th percentile of both ambient temperature and wind speed,
where 𝑝 is the pressure and 𝜇 is the dynamic viscosity. Together with throughout the year. These were 42.04 ◦ C and 5.6 m/s, respectively,
the continuity equation, the three equations would suffice to close and are used for the 𝑇𝑎𝑚𝑏 and 𝑈𝑤𝑖𝑛𝑑 boundary conditions in Fig. 2. The
a 2D fluid flow problem in the laminar regime. However, as stated, 90th percentile for irradiation is also used to establish the viability of
for the typical atmospheric wind flow over a typical PV module, the the proposed enhancement, since much of the power generation occurs
flow is predominantly turbulent. At the chosen flow conditions, it is at the higher end of irradiation. These three points coincide at 12:35
indeed fully turbulent. The 𝑘-𝜔 turbulence model was used to relate the pm local time on the 25th day of May. At these conditions, the flow is
additional terms in the equations [27,28]. Specifically, the 𝑘-𝜔 model fully turbulent.1
solves for the turbulent kinetic energy, 𝑘, and the specific dissipation With the recorded incident heat flux (pyranometer reading) and
rate, 𝜔, power output, the remainder of the heat flux transformed to thermal
[( ) ] energy at the surface of the cells (𝑞𝑠 ) can be calculated as,
𝜕(𝜌𝑢𝑗 𝑘) 𝜕 𝜌𝑘 𝜕𝑘
= 𝜌𝑃 − 𝛽 ∗ 𝜌𝜔𝑘 + 𝜇 + 𝜎𝑘 (4) Module Output Power
𝜕𝑥𝑗 𝜕𝑥𝑗 𝜔 𝜕𝑥𝑗 𝑞𝑠′′ = Incident Heat Flux − (8)
Area 𝑐𝑒𝑙𝑙𝑠,𝑡𝑜𝑡
𝜕(𝜌𝑢𝑗 𝜔) 𝛼𝜔
= 𝜌𝑃 − 𝛽𝜌𝜔2
𝜕𝑥𝑗 𝑘
[( ) ] 1
𝜕 𝜌𝑘 𝜕𝜔 𝜌𝜎 𝜕𝑘 𝜕𝜔 The complete data for that day is shown in Figs. A.15, A.16, and A.17 in
+ 𝜇 + 𝜎𝜔 + 𝑑 (5) the Appendix.
𝜕𝑥𝑗 𝜔 𝜕𝑥𝑗 𝜔 𝜕𝑥𝑗 𝜕𝑥𝑗

3
R.I. Bourisli et al. Solar Energy 269 (2024) 112345

Fig. 4. The design variables; the subject of the study.

2.4. Design variables and targets of the optimization

Guiding vanes of different shapes and sizes are affixed to an array


of PV modules. The locations, sizes, and radii of curvature of these
vanes are the subject of the optimization study. The goal is to find the
combination that yields the maximum reduction in the temperature of
the PV modules. Fig. 4 shows the various lengths considered as design
variables in the optimization study: the 𝑥- and 𝑦-coordinates of the
lower end (Start) of the vane, XS and YS, the 𝑥- and 𝑦-coordinates of
the upper end (End) of the vane, XE and YE, and the radius of curvature
(ROC) of the vane.

3. Model validation

Validation of the FE model is done in three stages. First, the fidelity


of the model in simulating turbulent flow is established. This is done
through comparison with published correlations for the local Nusselt
number for a simple turbulent flow over a flat plate. In conjunction
with the first stage, mesh sensitivity analysis is performed to determine
any required mesh refinement. Finally, the validity of the PV module
numerical model is established by simulating the basic case described
in Section 2—the case of the single PV module in the Kabd testing
setup. This is modeled to make sure the numerical solution matches the
physical reality. Mesh refinement is continued throughout this stage,
to guarantee adaption of the mesh to the physics of the flow. The
introduction of the guiding vanes and the subsequent optimization
study are done using the final, validated, and verified numerical model.

3.1. Turbulent flow over flat plate

The flow regime at the base of the module is determined using


(local) Reynolds number,
𝜌𝑈 𝑥
𝑅𝑒𝑥 = (9)
𝜇
where 𝑈 is the free-stream velocity and 𝑥 is the dimension along the
𝑥-axis (ground), which is the distance from inlet to the base of the
first module. This uniform inlet velocity at the given distance from
Fig. 3. PV module and data logging system use in the validation.
the module was found to be realistic, so that the turbulent boundary
layer envelops the entire first modules. The values corresponding to
the base of the PV module is 𝑥 = 1.5 m. The free-stream wind velocity
corresponding to the chosen time of day is 5.6 m/s. Other properties
where the area is that of nine 0.16 by 0.16 m cells, as shown in
are 𝜌 = 1.145 kg∕m3 and 𝜇 = 1.895(10−5 ) Pa s. This gives a Reynolds
Fig. 1(b). The flux, PV power output, thermocouple temperatures, and number of 5.44(106 ), which is in the fully-turbulent regime.
the average temperature recorded by the data logger at that instant are The critical location along the 𝑥-direction, where 𝑅𝑒𝑐𝑟 = 5(105 ),
given in Table 1. is calculated as 𝑥𝑐𝑟 = 1.38 m, which means that while the region of

4
R.I. Bourisli et al. Solar Energy 269 (2024) 112345

Table 1
Properties and boundary conditions raw values at the time instant for the validation process.
Day/Time Pyranometer Power Heat fluxa tc1 tc2 tc3 tc4 𝑇𝑃 𝑉 ,𝑎𝑣𝑒
May 25 (W/m2 ) (W) (W/m2 ) (◦ C) (◦ C) (◦ C) (◦ C) (◦ C)
12:34:48 948.34 32.73 806.27 56.72 55.46 54.53 55.25 55.49

Environmental 𝑈∞ 𝜌 𝜇 Re 𝑇𝑎𝑚𝑏
Conditions (m/s) kg/m3 (Pa s) (◦ C)
5.6 1.145 1.895(10−5 ) 5(10)5 42.04
a The bottom face heat flux due to the albedo effect was taken to be 15% of this top face value.

3.3. Model validation for a single PV module (without vanes)

With the validated turbulence model and the optimized mesh for
the basic case, the model was verified using the physical case from
the testing rig at Kabd, in the desert of Kuwait, Fig. 3. The single PV
module is modeled and the maximum temperature at the back surface
is compared to that of the physical model. The module dimensions,
irradiance, wind velocity, and ambient temperature given in Section 2.3
are used.
Temperature readings at 4 different locations on the back surface of
the module were recorded at 30-second intervals. On the specified day
Fig. 5. Comparison of the FE-computed heat transfer coefficient along the ground with and time (25th of May, at 12:35 pm), the 4 thermocouples readings
that obtained from the correlation of Eq. 3.4 in [30].
were averaged. The average temperature was 55.49 ◦ C. Results from
the FE model show that the average temperature over the back surface
for these conditions is 55.04 ◦ C, which is close enough to the physical
interest is in the fully turbulent regime, the plate upstream of the PV model. Fig. 6 in the next section shows the velocity streamlines and
module could have mixed laminar/turbulent flow over it. temperature contours for this validation case of a single PV module
The local heat transfer coefficient is calculated numerically via, with no vanes.
𝑞𝑥′′
ℎ𝑥 = (10) 4. Results and discussion
𝑇 − 𝑇∞
where 𝑞𝑥′′ is the incident solar irradiation, 𝑇∞ is the ambient dry bulb A sequential parametric study is performed for the optimum guiding
temperature, and 𝑇 is the local surface temperature along the length vanes which can be considered as arc segments of circles with given
of the surface. The Nusselt number correlation proposed by Bergman radii. A series of vanes with varying coordinates for their starting and
et al. [30] is used, ending point, as well as radii of curvature, were simulated. Bounds on
h 𝐿 ( 4∕5
) the values tested were determined by several physical arguments and
Nu𝐿 = 𝐿 = 0.037Re𝐿 − 𝐴 Pr1∕3 (11) exploratory numerical simulations. Additionally, from the preliminary
𝑘
0.6 ≤ Pr ≤ 60 (11a) runs, an ROC of 1.05 m was found to be best, across all guiding vanes
geometries, locations, and modules. The ROC value was thus fixed and
Re𝑥,𝑐 ≤ Re𝐿 ≤ 108 (11b) eliminated from the decision variables.
𝐴 = 0.037Re4∕5 1∕2
(11c) From these initial runs also came the effective, meaningful ranges
𝑥,𝑐 − 0.664Re𝑥,𝑐
of the other decision variables—the ones that determine the size, loca-
Fig. 5 shows the local heat transfer coefficient along the flat plate. tion and orientation of the guiding vanes. These ranges are XS(8.90–
The numerical (CFD) solution shows closeness to the result of the cor- 9.05 m), YS (0.08–0.14), XE (9.10–9.25), and YE (0.18–0.24), where X
relation. Furthermore, the simulated and calculated Nusselt numbers and Y denote the 𝑥- and 𝑦-coordinates and S and E denote the start and
were 23.83 and 23.14, respectively, with a difference of about 0.69, or end points of the arc; reference is made to Fig. 4. These numbers pertain
2.89%, which was deemed acceptable. to coordinates in the channel for the first module/vane. Adjustments
in XS and XE values are made for subsequent vanes according to the
3.2. Mesh sensitivity analysis distance from the first module.

The FE model was repeatedly run for gradually increasing mesh 4.1. Single PV module design (with/without vanes)
sizes and the influence on the results was observed. The maximum sur-
face temperatures in the FE model were noted and compared to those Results for the base case of a module without guiding vanes was
from the experimental model. Specifically, the maximum temperature presented in the validation phase of Section 3.3, where the maximum
of the lower surface of the PV module was used as the compari- temperature in the module was 59.26 ◦ C.2 Velocity streamlines (su-
perimposed on velocity magnitude contours) and temperature contours
son criteria. Table 2 shows (part of) the mesh, number of elements,
(with velocity vectors) are shown in Fig. 6. It can be seen from the
and maximum temperature for each mesh refinement level. Level 1
figures that a large eddy develops and attaches to the back surface
refinement was deemed acceptable because the change in maximum
temperature from this level to the next is less than 1 degree Celsius. It
is noted that the mesh concentrates elements around and downstream 2
The maximum temperature in the module is slightly higher (less than
of the module, in the wake, where fluid and energy recirculate back to 0.5 ◦ C) than the surface temperature at the back surface of the module, due to
the back of the module. However, because of the marginal additional the direct irradiation incident on the top surface. But since validation was done
computational cost associated with using the higher level mesh (Level using back surface temperature, these will be used to measure the effectiveness
2), all results reported used the Level 2 mesh. of the design.

5
R.I. Bourisli et al. Solar Energy 269 (2024) 112345

Table 2
Maximum temperature with mesh refinement level.
Refinement level Mesh shapea Number of elements Maximum temperature (◦ C) Difference (◦ C)

0 2,382 58.604 –

1 4,959 59.277 0.67

2 10,765 59.259 0.018

a
Only part of the mesh is shown. The downstream of the module is cut.

Table 3
Optimum results for the guiding vane for one PV module.

No. XS YS XE YE Temperature

1 8.90 0.12 9.10 0.24 50.72 C
2 8.90 0.10 9.10 0.22 50.93 ◦
C
3 8.90 0.08 9.10 0.2 51.09 ◦C

4 8.90 0.14 9.10 0.24 51.13 ◦C ⟸



5 8.95 0.12 9.10 0.22 51.13 C

6 8.95 0.12 9.10 0.24 51.14 C

7 9.00 0.08 9.10 0.2 51.25 C

8 9.00 0.14 9.10 0.24 51.25 C
9 9.05 0.14 9.10 0.22 51.31 ◦C

10 8.90 0.08 9.10 0.24 51.32 ◦C

of the module. The low-kinetic-energy, slow-moving recirculating air


in the eddy traps the heat from the module and results in the high
temperature observed, namely, 59.26 ◦ C in a 42.04 ◦ C environment.
Table 3 shows the shapes and locations for a number of the best-
performing vanes and the corresponding highest surface temperature in
the module. The temperature dropped from 59.26 ◦ C to around 51◦ C
when using the guiding vanes, a reduction of more than eight degrees
Celsius. Fig. 7 show the velocity streamlines and temperature contours,
respectively, for the ‘‘optimum’’ vane case. Compared to streamlines in
Fig. 7, it is clearly noticed that the flow in concentrated over the back
face of the module. As seen in Fig. 7(b), the large recirculation zone is
moved away from the module so that the energy trapped is not adjacent
Fig. 6. Fluid flow and temperature contours results for a single PV module; without
to the back of the module.
vanes (the validation case).
Even though the lowest temperature occurred for solution no. 1,
that vane, with its low vertical starting point (a YS of 0.12), leaves
little air for any downstream modules. Therefore, the vane selected
one line will inevitably cause a rise in the temperatures of downstream
for the first module and fixed for all subsequent analyses (the 2-, 3-
modules since the air used to cool it is now warmer.
and 4-module cases) will be the one resulting in the minimum temperature
among the ones with the highest starting points (YS), i.e., guiding vane no. Without guiding vanes in any of the two modules, the maximum
4 (highlighted). This was found to allow more air to flow past the first surface temperature in the short array is 68.07 ◦ C, which was located
module with sufficient momentum to enable cooling of the modules low in the second module, where wind grinds to a halt due to opposite
downstream while still having an appreciable temperature reduction eddies which result in a stagnation point. Fig. 8 shows the velocity
(𝑇𝑚𝑎𝑥 = 51.13◦ C, compared to 50.72 ◦ C for the absolute optimum vane). streamlines and temperature contours before adding vanes to the two
It is noticed that before adding the vane, the maximum temperature PV modules.
in the module was on the back surface of the PV module, hidden from The optimum configuration for the first guiding vane is applied
the main flow. After adding the vane, the maximum temperature shifted to the first module in the two-module array. The ranges employed
to the front side of the PV module, facing the flow. This phenomenon for the second vane, after a few exploratory runs, were also used for
demonstrates the effectiveness of the guiding vane in removing enough the second, albeit with a shift in the 𝑥-direction equal to the distance
energy from the back surface so as to shift the maximum temperature between the bottoms of the modules.
to the surface directly being irradiated. Table 4 shows the top 10 optimum results and the corresponding
maximum temperature. It is noted that the YS value for the second
4.2. Two PV module design module is lower than that of the first—0.10 vs. 0.14 m (or vs. 0.12 m).
The minimum temperature was 55.47 ◦ C, a nontrivial drop of 12.6 ◦ C
To further mimic the layout of a solar plant, a second PV module, compared to the case with no vanes. Fig. 9 shows the velocity stream-
with the same dimensions and angle, is added behind the first one, at a lines and temperature contours for the flow over two modules after
distance of 1 m between their lower ends. Having more PV modules in adding guiding vanes to both.

6
R.I. Bourisli et al. Solar Energy 269 (2024) 112345

Fig. 9. Fluid flow and temperature contours results for two PV modules; with vanes.

Table 4
Optimum results for the guiding vane of the second module.

No. XS YS XE YE Temperature
1 10.05 0.10 10.10 0.20 55.47 ◦ C⟸

2 10.05 0.10 10.10 0.18 56.38 ◦C



3 10.05 0.10 10.15 0.18 56.51 C

4 10.00 0.10 10.10 0.18 56.86 C
Fig. 7. Fluid flow and temperature contours results for a single PV module; with vanes ◦
5 10.05 0.10 10.15 0.20 57.12 C
(fourth best). ◦
6 10.00 0.10 10.10 0.20 57.16 C
7 10.05 0.12 10.10 0.22 57.33 ◦
C
8 10.00 0.10 10.15 0.20 57.85 ◦C

9 10.05 0.12 10.10 0.20 57.97 ◦C



10 9.95 0.10 10.10 0.20 58.07 C

Fig. 8. Fluid flow and temperature contours results for two PV modules; without vanes.
Fig. 10. Fluid flow and temperature contours results for three PV modules; without
vanes.

4.3. Three PV modules design (with/without vanes)

the first and second modules. A lower starting point for the vane allows
The base, no-vane results for the three-PV-module case was sim-
it to gather up more of the available air, which diminished after going
ulated and the maximum temperature of 69.06 ◦ C among the three
past two modules. The third guiding vane is optimized separately, after
modules was still in the second PV module. Fig. 10 shows the veloc- fixing the parameters of the first and second vanes to the optimum
ity streamlines and temperature contours. One reason for the higher values found previously. Table 5 shows the top optimum designs and
temperature being in the middle module is that the eddy behind the the corresponding maximum temperatures.
third and last module is exposed to more (cooler) ambient air. This is Fig. 11 shows the velocity streamlines and temperature contours for
unlike the second module, which is sandwiched between two modules. the flow over three modules after adding guiding vanes to all modules.
When optimizing the third guiding vane, its lower point is expected It can be seen that for this optimum case, the maximum temperature
to be even lower, since air lost more momentum after going through was 61.09 ◦ C, a significant, 8-degree-drop from 69.06 ◦ C. It is also

7
R.I. Bourisli et al. Solar Energy 269 (2024) 112345

Table 6
Optimum results for the guiding vane of the fourth module.

No. XS YS XE YE Temperature
1 12.00 0.03 12.10 0.13 62.8 ◦ C⟸
2 12.00 0.03 12.10 0.14 62.93 ◦ C
3 11.95 0.03 12.10 0.15 63.87 ◦ C
4 12.00 0.03 12.10 0.12 63.9 ◦ C
5 12.00 0.04 12.10 0.15 64.04 ◦ C
6 12.00 0.04 12.10 0.12 64.24 ◦ C
7 12.00 0.04 12.10 0.13 64.33 ◦ C
8 11.95 0.03 12.10 0.14 64.37 ◦ C
9 12.00 0.04 12.10 0.14 64.42 ◦ C
10 11.95 0.03 12.10 0.13 64.51 ◦ C

Fig. 11. Fluid flow and temperature contours results for three PV modules; with vanes.

Table 5
Optimum results for the guiding vane of the third module.

No. XS YS XE YE Temperature
1 11.05 0.06 11.10 0.15 61.09 ◦ C⟸

2 11.05 0.06 11.10 0.13 61.69 ◦C



3 11.05 0.06 11.15 0.15 61.83 C

4 11.00 0.07 11.10 0.19 61.84 C

5 11.00 0.07 11.10 0.18 62.33 C

6 11.00 0.06 11.10 0.15 62.35 C

7 11.00 0.07 11.10 0.16 62.70 C
◦C Fig. 13. Fluid flow and temperature contours results for four PV modules; with vanes.
8 11.00 0.07 11.10 0.16 62.70
9 11.00 0.07 11.10 0.17 62.98 ◦C

10 11.00 0.08 11.10 0.19 63.31 C
reduction in momentum, going from free stream conditions to the wake
behind the first module. Just as was done for the previous cases, the
worth noting that the location of the maximum temperature shifted guiding vane for the fourth PV module was optimized separately.
from the second module to near the top of the third. This makes more Table 6 shows some of the optimum designs obtained, with the
sense if we consider that the more air flows over the modules, the more corresponding maximum temperatures in the array. It can be seen that
it heats up. the minimum surface temperature reached was 62.8 ◦ C, constituting a
significant drop of more than 9 degrees from 72.05 ◦ C, for the case of no
4.4. Four PV modules design (with/without vanes) vanes. Fig. 13 shows the velocity streamlines and temperature contours
around the four PV modules for the optimum case. Similar to the
three-module case, the location of the maximum temperature moved
[t] Finally, a fourth PV module is added and its vane is optimized.
from the second to the fourth module, signifying the effectiveness of
In the case of four PV modules in a row, with no guiding vanes, the
the resulting flow pattern in distributing the load (of carrying away
maximum surface temperature occurred, again, in the second module;
the thermal energy) among all vanes. This provides evidence that the
it was 72.05 ◦ C. Fig. 12 show the velocity streamlines and temperature
results obtained indeed tend towards the global optimum.
contours around the modules before adding vanes to any module.
Similar to previous cases, much of the action takes place behind the first
4.5. Air flow distribution and efficiency
module—cf. Fig. 12(a). This is where the flow experiences the largest

When it comes to the breakdown of air quantities going behind each


module, Fig. 14 shows the fractions of the air mass flow rates going
above and below each of the four guiding vanes. It can be seen that
the flow is well-distributed among each of the modules. The goal of this
work is to have this air, the cooling fluid, optimally distributed among
the heat sources. The results presented, in addition to the fractions seen
in Fig. 14, suggest that this was accomplished.
The efficiency of PV modules was the ultimate goal of the study. A
typical silicon PV cell is usually rated at STC, an operating temperature
of 25 ◦ C. As stated, the general rule-of-thumb is that for every 1-
degree Celsius increase in temperature, a 0.3%-drop in efficiency is
experienced [2],

𝜂𝑎𝑐𝑡 = 𝜂𝑛𝑜𝑚𝑖𝑛𝑎𝑙 − (𝑇𝑎𝑐𝑡 − 25)(0.3∕100)𝜂𝑛𝑜𝑚𝑖𝑛𝑎𝑙 (12)

For example, the 2-module array studied here, which might ideally
Fig. 12. Fluid flow and temperature contours results for four PV modules; without have a 20%-efficiency if operating at STC, will have an actual effi-
vanes. ciency of 17.4% when operating at 68.07 ◦ C. However, with the use

8
R.I. Bourisli et al. Solar Energy 269 (2024) 112345

Fig. 14. Air mass flow rate distributions above and below the guiding vanes for the case of four PV modules.

Table 7
Summary of optimum guiding vanes dimensions and locations, the maximum temperatures on the back surface of the modules before and after
the addition of the vanes, and the reduction in these temperatures.
PV modules Optimum 𝑇𝑚𝑎𝑥 (◦ C) 𝑇𝑚𝑎𝑥 reduction
XS YS XE YE w/o Vanes w/ Vanes
Single 8.90 00.14 9.10 00.24 59.26 51.13 8.13
Double 10.05 00.10 10.10 00.20 68.07 55.47 12.6
Triple 11.05 00.06 11.10 00.15 69.06 61.09 8.00
Quadruple 12.00 00.03 12.10 00.13 72.05 62.8 9.25

of optimized guiding vanes such as the ones proposed here, and if back surfaces of the modules, where the highest temperatures usually
the temperature is reduced from 68.07 ◦ C to 55.47 ◦ C as predicted, occur. The reduction in the maximum temperature of the modules
the new module efficiency becomes 18.2%, nearly 0.8%-increase in ranged from 8 to 12.6 ◦ C, for the different situations. The resulting set
absolute efficiency. This kind of improvement, when implemented on a of vanes distribute the air (cooling potential) proportionately among
large scale, translates into huge power generation potential. All this for the arrays, from 28% for the first of four to 73% for the fourth of the
the cost of the passively-operating, readily-optimized, easily-installed four, as seen in Fig. 14. Furthermore, they altered the airflow over the
guiding vanes. Passive cooling techniques hold a sustained advantage back surfaces of the modules significantly such that thermal dead zones
over active one by not involving a power source, moving parts, nor are considerably reduced or eliminated altogether. It is these dead
separate, intricate mechanical/electrical systems. zones that cause heat build-ups and temperature rises by preventing
As a final summary, Table 7 gives the optimum guiding vanes di- the heat from advected away. The flow field is altered in a minimally
mensions, locations, and the resulting reductions in maximum temper- invasive fashion, using these simple to design, manufacture and install
ature they lead to, for the four cases tested. The maximal temperatures vanes.
reported are for the bottom surfaces of the modules since the validation The height above the ground plays an important role in determining
was done using data taken on the lower surfaces. As stated before, the the amount of air that can flow past the first module too. In this analy-
difference between the temperatures of the top and bottom surfaces of sis, a maximum distance of 0.14 m was assumed, based on experience
the cells is very small due to the negligible resistance to conduction and observations. This can be the subject of further investigation.
heat transfer in the solid, compared to that to convection.3 Simulations Furthermore, the current analysis was limited to PV modules facing
were repeated for the overall maximum temperature that considers south, with a 30◦ inclination angle. While this is the rule of thumb
all faces and all modules, with and without vanes, and the resulting for Kuwait, further analyses can also allow for different azimuth angles
temperature drops were 7.57, 14.25, 9.37, and 13.52 ◦ C, compared to if they can help increase the total power generation. In addition, the
8.13, 12.6, 8.00, and 9.25 ◦ C, which are quite close. Simulations were effect of the radius of the arc shape guiding vane can be studied further.
also repeated considering the maximum temperatures within the entire Lastly, the study does not include the effects of soiling or shading of the
modules, not just the surface temperatures, and the results, temperature PV modules.
reductions and optimum dimensions were indistinguishable from the
ones reported here, albeit a few degrees higher. Declaration of competing interest

5. Conclusions and recommendations The authors declare that they have no known competing finan-
cial interests or personal relationships that could have appeared to
In this study, guiding vanes that force ambient air to flow behind influence the work reported in this paper.
PV modules are proposed. The objective of the study was to reduce the
temperatures of the PV module, which will enhance their efficiency. Acknowledgement
The shapes, sizes, and locations of the guiding vanes were sequentially
optimized for 1, 2, 3 and ultimately 4 inline PV modules. An FE model The authors would like to acknowledge the partial support of this
is developed and validated using real data (power, irradiance, wind work by Kuwait Foundation for the Advancement of Sciences (KFAS),
speed, and ambient temperature) from the outdoor test setup in Kabd, grant no. CN1815EE01.
Kuwait.
The existence of the vanes truly helped change the momentum of
Appendix. Average real time data for May 25
the air flowing under the module and directed it upward and over the
Solar testing rig measurements were taken on 25 May and com-
3
It should also be mentioned that initial runs involved modeling all layers
pared to typical meteorological year (TMY) data. Graphs Fig. A.15,
of the cell, including glass, encapsulant and silicon layers. Some layers were Fig. A.16 and Fig. A.17 show the recorded data for the solar radiation
a couple hundred micron thick, which added a huge computational cost for (W/m2 ), ambient temperature (◦ C), and wind speed (m/s), respectively,
an unnoticeable gain in accuracy. The module was thus taken to be one solid throughout the day. A polynomial best-fit line is also shown (in red) for
surface, with predominantly glass properties. each quantity: quartic for heat flux, sextic for ambient temperature, and

9
R.I. Bourisli et al. Solar Energy 269 (2024) 112345

quartic for wind speed. The equations for these lines are shown below,
where 𝑡 is in hours.

𝑞𝑠′′ = 0.3755𝑡4 − 17.873𝑡3 + 283.42𝑡2


− 1686.5𝑡 + 3355.7 (A.1)
−5 6 5 4
𝑇𝑎𝑚𝑏 = −1 × 10 𝑡 + 0.0009𝑡 − 0.0241𝑡
+ 0.2508𝑡3 − 0.8186𝑡2 + 0.1189𝑡 + 30.454 (A.2)
𝑉𝑤𝑖𝑛𝑑 = −4.7 × 10−3 𝑡3 + 0.1606𝑡2 − 1.3518𝑡
+ 6.2318 (A.3)

References

[1] J.K. Kaldellis, M. Kapsali, K.A. Kavadias, Temperature and wind speed im-
pact on the efficiency of PV installations. Experience obtained from outdoor
measurements in Greece, Renew. Energy 66 (2014) 612–624.
[2] M. Hasanuzzaman, A.B.M.A. Malek, M.M. Islam, A.K. Pandey, N.A. Rahim, Global
advancement of cooling technologies for PV systems: A review, Sol. Energy 137
(2016) 25–45.
Fig. A.15. Global incident plane of array radiation (heat flux) at the surface of the [3] P. Dwivedi, K. Sudhakar, A. Soni, E. Solomin, I. Kirpichnikova, Advanced cooling
module vs. time-of-day for the 25th of May. techniques of P.V. modules: A state of art, Case Stud. Therm. Eng. 21 (2020)
100674.
[4] S. Rahmanian, H.R. Koushkaki, A. Shahsavar, Numerical assessment on the
hydrothermal behaviour and entropy generation characteristics of boehmite
alumina nanofluid flow through a concentrating photovoltaic/thermal system
considering various shapes for nanoparticle, Sustain. Energy Technol. Assess. 52
(B) (2022) 102143.
[5] M. Jaszczur, J. Teneta, Q. Hassan, E. Majewska, R. Hanus, An experimental
and numerical investigation of photovoltaic module temperature under varying
environmental conditions, Heat Transf. Eng. 42 (3–4) (2021) 354–367.
[6] I.M. Kirpichnikova, K. Sudhakar, I.B. Makhsumov, A.S. Martyanov, S.S. Priya,
Thermal model of a photovoltaic module with heat-protective film, Case Stud.
Therm. Eng. 30 (2022) 101744.
[7] D. Sato, N. Yamada, Review of photovoltaic module cooling methods and
performance evaluation of the radiative cooling method, Renew. Sustain. Energy
Rev. 104 (2019) 151–166.
[8] 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 (1) (2009) 91–96.
[9] Z.A. Haidar, J. Orfi, Z. Kaneesamkandi, Experimental investigation of evaporative
cooling for enhancing photovoltaic panels efficiency, Results Phys. 11 (2018)
690–697.
[10] A.H.A. Salman, K.H. Hilal, S.A. Ghadhban, Enhancing performance of PV module
using water flow through porous media, Case Stud. Therm. Eng. 34 (2022)
102000.
Fig. A.16. Ambient temperature at the location of the module vs. time-of-day for the [11] M.J. Huang, P.C. Eames, B. Norton, Phase change materials for limiting tem-
25th of May. perature rise in building integrated photovoltaics, Sol. Energy 80 (9) (2006)
1121–1130.
[12] M.J. Huang, The effect of using two PCMs on the thermal regulation performance
of BIPV systems, Sol. Energy Mater. Sol. Cells 95 (3) (2011) 957–963.
[13] A. Hasan, S.J. McCormack, M.J. Huang, B. Norton, Evaluation of phase
change materials for thermal regulation enhancement of building integrated
photovoltaics, Sol. Energy 84 (9) (2010) 1601–1612.
[14] E. Skoplaki, J.A. Palyvos, Operating temperature of photovoltaic modules: A
survey of pertinent correlations, Renew. Energy 34 (1) (2009) 23–29.
[15] F. Fouladi, P. Henshaw, D.S.-K. Ting, S. Ray, Wind turbulence impact on solar
energy harvesting, Heat Transf. Eng. 41 (5) (2019) 407–417.
[16] H.G. Teo, P.S. Lee, M.N.A. Hawlader, Wind turbulence impact on solar energy
harvesting, Appl. Energy 90 (1) (2012) 309–315.
[17] B.J. Brinkworth, B.M. Cross, R.H. Marshall, H. Yang, Thermal regulation of
photovoltaic cladding, Sol. Energy 61 (3) (1997) 169–178.
[18] B.J. Brinkworth, Thermal regulation of photovoltaic cladding, Sol. Energy 80 (9)
(2006) 1131–1134.
[19] M.S. Sheik, P. Kakati, D. Dandotiya, M.U. Ravi, C.S. Ramesh, A comprehensive
review on various cooling techniques to decrease an operating temperature of
solar photovoltaic panels, Energy Nexus 8 (2022) 100161.
[20] A.M. Elbreki, A.F. Muftah, K. Sopian, H. Jarimi, A. Fazlizan, A. Ibrahim,
Experimental and economic analysis of passive cooling PV module using fins
and planar reflector, Case Stud. Therm. Eng. 23 (2021) 100801.
[21] J. Li, Y. Zhou, X. Niu, S. Sun, L. Xu, Y. Jian, Q. Cheng, Performance evaluation
of bifacial PV modules using high thermal conductivity fins, Sol. Energy 245
(2022) 108–119.
Fig. A.17. Wind speed at the location of the module vs. time-of-day for the 25th of [22] A. Royne, C.J. Dey, Design of a jet impingement cooling device for densely
May. packed PV cells under high concentration, Sol. Energy 81 (8) (2007) 1014–1024.
[23] P. Valeh-e-Sheydaand, M. Rahimi, A. Parsamoghadam, M.M. Masahi, Using a
wind-driven ventilator to enhance a photovoltaic cell power generation, Energy
Build. 73 (2014) 115–119.

10
R.I. Bourisli et al. Solar Energy 269 (2024) 112345

[24] A. Almuwailhi, O. Zeitoun, Investigating the cooling of solar photovoltaic [27] D.C. Wilcox, Basic Fluid Mechanics, second ed., DCW Industries, La Cañada,
modules under the conditions of Riyadh, J. King Saud Univ., Eng. Sci. 35 (2) 1998.
(2023) 123–136. [28] D.C. Wilcox, Formulation of the k-w turbulence model revisited, AIAA J. 46 (11)
[25] Y. Zhang, C. Shen, C. Zhang, J. Pu, Q. Yang, C. Sun, A novel porous channel (2008) 2823–2838.
to optimize the cooling performance of PV modules, Energy Built Environ. 3 (2) [29] COMSOL Multiphysics® Reference Manual, v. 6.0, COMSOL AB, Inc., Stockholm,
(2022) 210–225. Sweden, 2019.
[26] Y.A. Çengel, J.M. Cimbala, Fluid Mechanics: Fundamentals and Applications, [30] T.L. Bergman, A.S. Lavine, F.P. Incropera, D.P. DeWitt, Fundamentals of Heat
fourth ed., McGraw Hill, 2017. and Mass Transfer, seventh ed., John Wiley & Sons, 2011.

11

You might also like