You are on page 1of 29

Journal Pre-proofs

Thermal performance of a discontinuous finned heatsink profile for PV pas-


sive cooling

J.G. Hernandez-Perez, J.G. Carrillo, A. Bassam, M. Flota-Banuelos, L.D.


Patino-Lopez

PII: S1359-4311(20)33717-0
DOI: https://doi.org/10.1016/j.applthermaleng.2020.116238
Reference: ATE 116238

To appear in: Applied Thermal Engineering

Received Date: 15 April 2020


Revised Date: 16 September 2020
Accepted Date: 16 October 2020

Please cite this article as: J.G. Hernandez-Perez, J.G. Carrillo, A. Bassam, M. Flota-Banuelos, L.D. Patino-Lopez,
Thermal performance of a discontinuous finned heatsink profile for PV passive cooling, Applied Thermal
Engineering (2020), doi: https://doi.org/10.1016/j.applthermaleng.2020.116238

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier Ltd.


Thermal performance of a discontinuous finned heatsink profile
for PV passive cooling
J.G. Hernandez-Perez1, J.G. Carrillo2*, A. Bassam3, M. Flota-Banuelos3, L.D. Patino-
Lopez1
1CONACYT - Centro de Investigacion Cientifica de Yucatan, Renewable Energy Department, Parque
Cientifico y Tecnologico de Yucatan, C.P. 97302, Mérida, México.
2Centro de Investigacion Cientifica de Yucatan, Materials Department, C. 43, No. 130, Chuburna de Hgo.,
C.P. 97205, Mérida, México.
3Universidad Autonoma de Yucatan, Engineering Faculty, Industrias no contaminantes, A.P. 150, Merida,
Mexico.

*Corresponding author: jgcb00@cicy.mx

ABSTRACT

In this work, an improved CFD model based on previous results is used to evaluate the
scaling effects on thermal and hydraulic performance of a passive heatsink design for PV
panel cooling purposes. PV systems efficiency degrades when operating temperature
increases, a common incidence in hot climates exposed to high solar irradiance. In order
to reduce PV modules operating temperature and maintain efficiency, a segmented fin
heatsink is proposed, whose performance under varying wind attack conditions
outperforms the conventional continuous fin profile heatsink design. Numerical
simulations using computational fluid dynamics software are presented, assessing
different levels of heatsink fin segmentation (fin width) and its effect on hydraulic
performance, as well as temperature level and homogeneity of the PV module. Two
prototypes were built for experimental evaluation: the model with better results in the
numerical tests, and a conventional design homologue, both manufactured following a
cost-effective ratio and using 22-gauge aluminum sheet. Numerical simulation showed a
temperature reduction of up to 7 °C using the proposed heatsink, as well as a lower
temperature gradient. Regarding hydraulic performance, the segmented fin profile
reduces pressure losses up to 49% when compared to a conventional heatsink.
Experimental results are in good agreement, as thermal images showed a mean
temperature reduction of around 5 °C and similar homogeneity. The proposed heatsink

1
design shows improved hydraulic characteristics, thus enhancing heat transfer using the
same amount of material as a conventional design.

Keywords: Passive cooling; Computational fluid dynamics; Finned heatsink; Heat transfer
enhancement; PV system

1. Introduction

PV Module performance strongly depends on environmental factors, such as:


ambient temperature, solar irradiance, wind speed and direction [1]. A portion of the solar
radiation absorbed by the PV cells is transformed into thermal energy, causing a
temperature increase. It has been widely reported that temperature is the major factor
that negatively affects the efficiency of photovoltaic systems based on crystalline silicon
(c-Si), which has the largest market share [2]. Researchers have determined that
efficiency losses caused by the rise in operational temperature of the PV modules can be
quantified at an average of 0.25%/°C and up to 0.5%/°C [3]. This temperature increase
produces a reduction in both the open circuit voltage and fill factor, as well as an increase
in the recombination of internal charge carriers, resulting in a decrease in the output
power of the PV system [4-8]. At the same time, it has also been found that thermal cycles
accelerate the aging of the system in the long term and degrade its performance [9, 10];
and, in some cases, sudden failure may occur due to cyclic thermal stress, for example
[11]: cracking of a cell, interconnection fatigue producing an open circuit, insulation
degradation producing a short circuit, top glass breakage, material delamination due to
differential expansion, increase in hot spot criticality, and bypass diode failure. Cochard
et al. [11] mentioned that according to the US military electronic reliability prediction
standard, the failure rate of microcircuits generically similar to PV cells increases 50%
when the component’s average junction temperature rises from 70 °C to 80 °C.
Furthermore, the authors indicate that manufacturers specify a normal application range
(which is in practice subjective) that may involve conditions that are unexpected and more
benign than some customers may require. An example is given in which [12] describe
how the adhesion between the backsheet and the PV cell encapsulant, Ethylene-Vinyl-

2
Acetate (EVA), of one module type was qualified by a damp heat test run at 65 °C instead
of 85 °C when the latter is the standard IEC 61215 and IEC 61646 PV qualification test
level. The manufacturer found that delamination rates increased non-linearly at some
temperature between 65 °C and 85 °C. The manufacturer concluded that modules used
in their specified conditions should not exceed 65 °C. As a consequence, the higher
temperatures associated with non-linear delamination rates should not be used to
accelerate module nominal ageing processes either. The authors conclude that product
acquisitions for temperature environments at the upper end of the general range should
pay attention to non-standard levels of qualification details.
Lowering the surface temperature of a PV module can reduce the thermal rate of its
degradation. This is done by cooling the system (or preventing it from heating) during
operation [13,14]. From an energy perspective, cooling systems can be classified as
active or passive. Active systems have been proven to achieve good results for PV
cooling [15-20]. However, they rely on an external device to enhance heat transfer by
keeping a working fluid in circulation; the need for such devices results in extra energy
consumption and higher operational costs, mainly due to the need for maintenance.
Passive cooling systems are a good low-cost alternative for PV thermal management,
as they operate without consuming energy and work based on three basic mechanisms
of heat transfer: convection, conduction, and radiation [21]. Previous authors conclude
that using the naturally circulating air as a working fluid is the simplest method to cool a
PV module. Some passive heat extraction methods include heat pipes, heat exchangers
(thermosyphon effect) and finned heatsinks [22-26]. Finned heatsinks are mainly used in
electronic and industrial applications, where operating conditions are known and, in some
cases, there are mechanical devices such as fans to increase convection. However, its
efficient use in PV systems faces certain challenges, mainly the unpredictable nature of
on-field wind flow for which a conventional heatsink geometry is not optimal (continuous
straight fins). On the other hand, the size of a commercial PV module makes it necessary
to opt for alternative manufacturing techniques that allow the production of a low cost and
low weight system. As this paper deals with the design and construction of a passive
finned heatsink, the literature review is focused on some of the more recent and relevant
contributions on that field, as summarized in Table 1.

3
Table 1
Works on finned-heatsink passive cooling for PV modules.
Work ref. Theoretical Experimental Heatsink design Temperature Efficiency
reduction
Ahmad et al. No Yes Finned aluminum Average of 6.1 °C Increase of 1.75%
[27] plate attached with on PV surface in electrical
thermal grease conversion
Fatih et al. Yes Yes Various aluminum 3.39 °C on PV back Power increase of
[28] fin arrays attached surface (highest 9.40 W (highest
with thermal paste recorded) recorded)
Grubisic et al. Yes Yes Aluminum ribs Average of 8 °C on Average increase of
[29] attached with PV back surface 5% in electrical
thermal glue (modeled) conversion
Ayman et al. Yes Yes Aluminum fins with From 4 to 5 °C on Increase of 0.9% in
[30] absorber plate cell surface electrical conversion
(highest observed)
Fatih et al. Yes Yes Porous aluminum From 0.22 to 0.54 Power increase of
[31] fins (metal foam) °C on PV back 7.26 W (highest
surface recorded)
Senthilkumar No Yes Aluminum heat 5.9 °C on PV back Increase in
[22] spreader, headers surface electrical yield of
and cotton wick 14%
Popovici et al. Yes No Aluminum Average of 10 °C on Increase of power
[23] perforated ribs PV cell output of up to
7.55%
Grubisic et al. No Yes Perforated Up to approximately Increase of 5% in
[32] aluminum fins 3 °C on PV back power yield
surface

Major finned passive heatsink designs found in the industry include [33-37]: stepped
fins, angled and vertical fins, cut channel, and cross cut. These designs are optimized in
many cases to operate in confined spaces and specific operating conditions, as well as
natural convection, forced convection, or a combination of both. However, due to costs
and weight, their use is mainly limited to applications in the cooling of electronic devices
that require heatsinks of reduced size (<100 mm x 100 mm x 30 mm), except in particular
cases such as power modules and Insulated Gate Bipolar Transistors (IGBTs). Fig. 1
shows some of the designs commonly used in the electronics industry. Fig. 1a shows the
straight vertical fin heatsink, the most well-known geometry and referred to in this present
work as "conventional heatsink". The heatsink with angled fins (Fig. 1b) improves the
conventional design by augmenting the dissipation surface by increasing the length of the
fins while maintaining the overall height of the heatsink.

Fig. 1.

4
Based on the literature review, the authors consider finned systems a viable
alternative for cooling a photovoltaic module, especially when cost, environmental,
reliability and ease of maintenance aspects are considered. It is also noteworthy that low-
cost systems with alternative geometries are generally proposed, resulting from
experimental tests while evaluating various designs. Some of these designs consist of
randomly positioned individual fins, which in some ways favors air flow when compared
to a conventional continuous channel heatsink. In [29] it is mentioned (and verified with a
numerical model) that a continuous fin heatsink can block the airflow depending on its
direction of incidence, i.e., a flow perpendicular to the fin channel. However, there is an
area of opportunity in optimizing heatsink parameters using numerical modeling as a
preliminary step to an experimental evaluation, which can lead to substantial savings in
time and money by discarding the fabrication and instrumentation setup of the worst
performing heatsink models. Additionally, specific hydraulic performance data are
obtained which help to indicate possible areas of improvement in heatsink design, i.e.,
which parameters are worth studying.
This paper presents a passive heatsink design that includes the advantages of using
a discontinuous fin profile, individual and angled fin profile. Angled fins allow a relatively
more compact system, where segmented profiles increase the area of heat dissipation,
which also favors heat extraction in conditions of natural convection with variable wind
speed and flow direction, reducing flow blockage and increasing turbulence. In their
previous work, Hernandez et al. [40] proposed a new fin array for PV passive heatsinks.
A segment of the PV module (unitary PV cell) was numerically modeled and the effect
of fin length and wind flow direction on thermal response was studied. It was found that a
discontinuous fin profile reduces pressure drop in the channels of the heatsink when
compared to a conventional geometry, thus enhancing convective heat transfer. In this
work, a full size heatsink is studied, given an increase in channel length results in a
negative effect on pressure drop. This has a direct impact on thermal and hydraulic
performance. It is also analyzed whether reducing the fin width (thus allowing a less
restrictive airflow) is a viable approach to further pressure loss reduction. First, a
numerical model of the PV module used in experimental tests as well as a set of heatsinks

5
was made. Three fin widths were evaluated for the discontinuous finned profile heatsink
and compared to a conventional heatsink, all regarding hydraulic and thermal
performance, in terms of pressure losses and PV module temperature level, respectively.
From numerical results, the discontinuous finned profile heatsink that reduced the
temperature of the PV module to a greater extent (as well as a conventional design
counterpart) was manufactured. Subsequently, an experimental evaluation was carried
out in a solar simulator to validate the numerical results. The findings of this work prove
that there is a possibility for improvement in passive heat dissipation applications for PV
modules. Modifying the design criteria to consider the constantly changing wind direction
in which the system operates allows further improvement in heatsink performance, thus
reducing PV module efficiency losses.

2. Materials and methods

2.1. Computational methodology

2.1.1. Governing equations

Conjugate heat transfer in an air-cooled PV module involves multiple solid and fluid
domains. The fluid domain includes the working fluid surrounding the PV module +
heatsink assembly, which is also within the heatsink channels. The physical aspects of
the fluids are governed by the following principles: conservation of the mass, conservation
of the moment and conservation of energy. These principles can be expressed as a set
of mathematical expressions called Navier-Stokes equations:
Continuity:

∂𝜌
∂𝑡 + ∇ ⋅ (𝜌𝑣) = 0 (1)

Moment:


( )
∂𝑡 𝜌𝑣 = ― ∇𝑝 + ∇ ∙ (𝜏) +𝜌𝑔 (2)

6
Energy:


∂𝑡(𝜌𝐸) + ∇ ⋅ (𝑣(𝜌𝐸 + 𝑝)) = ― ∇ ⋅ (∑𝑗ℎ𝑗𝐽𝑗) (3)

where 𝜌 is density, 𝑝 static pressure, 𝜏 is the stress tensor, 𝑣 is velocity, 𝐸 denotes


energy, ℎ is the species enthalpy, 𝐽 and is the diffusion flux. The governing equations are
discretized using the standard method of finite volumes to form a system of algebraic
equations which is solved by an iterative process using Computer Fluid Dynamics
software.

2.1.2. Turbulence model and near wall treatment

The RNG based k-e model used in this work is derived from the instantaneous Navier-
Stokes equations, using a mathematical technique called the “renormalization group”
method. RNG theory provides an analytical formula for turbulent Prandtl numbers, while
also including the effect of swirl turbulence as well as an additional term in its epsilon
equation to improve accuracy for rapidly strained flows, thus making the RNG k-e model
more accurate and reliable for a wider class of flows than the standard k-e model [41]. Its
turbulent kinetic energy and its dissipation rate are given by:


∂𝑡(𝜌𝑘)
∂ ∂
(
+ ∂𝑥𝑖(𝜌𝑘𝑢𝑖) = ∂𝑥𝑗 𝛼𝑘𝜇𝑒𝑓𝑓∂𝑥𝑗 + 𝐺𝑘 ― 𝜌𝜀
∂𝑘
) (4)

𝜀2

∂𝑡(𝜌𝜀) +

∂𝑥𝑖(𝜌𝜀𝑢𝑖) =

∂𝑥𝑗 ( ∂𝜀
𝛼𝜀𝜇𝑒𝑓𝑓∂𝑥𝑗 )+ 𝜀
𝐶1𝜀𝑘(𝐺𝑘) ― 𝐶2𝜀𝜌 𝑘 ― 𝑅𝜀 (5)

where 𝐺𝑘 represents the generation of turbulence kinetic energy due to the mean

velocity gradients, 𝛼𝑘 and 𝛼𝜀 are the inverse effective Prandtl numbers for 𝑘 and 𝜀,

respectively. Model constants 𝐶1𝜀 = 1.42 and 𝐶2𝜀 = 1.68 have values derived
analytically by RNG theory, where 𝑅𝜀 is the modification which improves the accuracy for

7
rapidly strained flows, and, 𝜇𝑒𝑓𝑓 is the effective viscosity derived from an analytical
differential equation that accounts for low-Reynolds number effects.
For this work, the enhanced wall treatment approach was chosen to resolve the
viscous sublayer of the flow. This near-wall modeling method combines a two-layer model
with enhanced wall functions. In this method, the viscosity affected near-wall region is
completely resolved all the way to the viscous sublayer. It uses one equation relationship
to evaluate the laminar sublayer with fine mesh and transition to log-low function for the
turbulent part of the boundary layer. The use of y+ with enhanced wall treatment provides
the scalable advantages over the standard wall functions; however, the restriction that
the near-wall mesh must be sufficiently fine everywhere can encumber computation.

2.1.3. Simulation setup and mesh configuration

The computational domain comprises a fluid limit box of 340 mm x 350 mm x 100 mm
which encloses a solar panel + heatsink (see Fig. 2a). Vertical walls of this enclosure are
extended 1 cm beyond the PV module edges and are defined as inlets or outlets
depending on the airflow direction of the study. This allows a suitable compromise
between accuracy and use of computational resources to model the conjugated heat
transfer phenomena. For ease of meshing, the model was divided in eight domains (see
Fig. 2b), which also have the properties of the standard layers of a PV module (Table 2).
Due to the geometric complexity of the fluid domain, it could not be fully discretized with
hexahedral mesh elements. In order to reduce the use of computational resources, it was
decided to divide this domain into two main volumes as shown in Fig. 2c. One was an
upper volume consisting of a structured mesh of hexahedral elements to resolve the
airflow on the upper surface of the PV module, and the other a lower volume that models
the airflow through the fins of the heatsink and backside of the PV module which was
discretized using tetrahedral elements.

Fig. 2.

8
As mentioned previously, the mesh was modeled to solve the viscous sublayer.
However, this imposes a high computational grid density. During the mesh dependence
study, it was found that a y+ value of around 3 (cell centroid inside the viscous sublayer)
yielded a good balance between accuracy and computing time. Keeping a constant y+
value around the whole geometry is difficult. This proved to be the case for the complex
heatsink geometry, where the value ranged from 3 to 5, and in some spots went up to 8.
Nonetheless, this was not deemed critical as flow separation is expected within the
heatsink, as well as a high turbulent intensity, in contrast with the upper PV wall, where
the air flows undisturbed. With the used mesh refinement, the number of grid elements
went from 4 x 106 for the reference PV module (no heatsink), to 13.9 x 106 for the
assembly with the most segmented heatsink. The convergence criterion to stop the
simulation was 1x10-6 for the scaled residual of the energy equation, and 1x10-4 for the
scaled residuals of the rest of the equations. The boundary conditions and model
parameters used are shown in Table 3.

Table 2
Thermo-physical properties of the model domains [42].
Domain Thickness (mm) Thermal Density (kg/m3) Specífic heat
conductivity (J/kg°C)
(W/m K)
Low-iron glass 3 1.8 3000 500
Cell 0.3 148 2330 677
EVA 0.5 0.35 960 2090
Tedlar 0.1 0.2 1200 1250
Aluminum 0.71 202.4 2719 871
Air @28 °C 100 0.026 1.171 1.006

Table 3
Boundary conditions, PV module characteristics and heatsink parameters.
Environment
Boundary condition Value Unit
Solar irradiance 1000 W/m2
Ambient 28 °C
temperature
Wind speed 2 m/s
Wind direction 0, 22.5, 45, 67.5, 90 °
PV module
Boundary condition Value Unit
Front Surface 0.91
emissivity [41]
Heatsink

9
Geometric Value Unit
parameter
Fin length 40 mm
Fin width 320 (1), 20 (16), 13.3 mm
(fin number per (24), 10 (32)
row)
Fin spacing 42 mm
Sheet gauge 22 (0.71) mm
(fin thickness)
Fin angle ±30° with Y axis

Several assumptions were made before performing numerical simulations:

● The problem is considered in 3D and steady state.


● All materials in the PV module have isotropic properties.
● The upper surface of the panel is exposed to a constant and homogeneous flow
of solar radiation.
● Ambient temperature is constant and equal on all sides of the PV module. Its
value corresponds to a three-year historical average for sunlight hours.
● Wind speed is constant at the model inlet, fluid flow disturbances are a direct
result of the PV module + heatsink assembly. The value of 2 m/s was chosen,
given it is a typical value of the studied region.

2.2. Numerical data analysis

A heat sink with optimal parameters is the product from this analysis. Three main
criteria are considered regarding heatsink performance: mean temperature level of the
PV module, PV temperature distribution, and heatsink pressure drop.
The PV module average surface temperature can be obtained by the software
through integration (as well as other variables on any other surface/volume on the model).
Temperature is directly related to efficiency, so the heatsink that achieves the lowest
surface temperature of the PV module also achieves the highest electrical conversion
efficiency.
The temperature distribution consists of surface contours which make it possible to
observe if a heatsink is prone to generating heat trapping/hot spots. Additionally, the

10
maximum and minimum surface temperature values can be obtained. A lower thermal
gradient is desirable, since large temperature differences can cause dispersion of PV
module electrical parameters and degrade operational efficiency.
The pressure drop is calculated by delimiting regions at the inlets and outlets of the
model in the lower fluid domain where the heatsink is located and obtaining the pressure
differential in these. Also, contours can be displayed showing an appreciation of critical
areas where pressure losses are abrupt, as well as the presence of eddies if a velocity
variable is shown (particle tracking or streamline).

2.3. Experimental methodology


2.3.1. Passive heatsink

Two 1100 aluminum alloy heatsinks were fabricated from commercially available 22-
gauge (0.71 mm) aluminum sheet, which is die cut and bent to form the fin profiles. Both
heatsinks had a 40 mm fin length and use the same amount of material. The conventional
variant (see Fig. 3a) has straight fins with a continuous profile and a 32 cm width. In
contrast, the proposed model obtained from a previous numerical design has fins with a
segmented profile every 2 cm (16 fins per profile) with an interleaved inclination of ±30°
with respect to the vertical axis (see Fig. 3b). This angled profile is designed to perform
better under mixed conditions owing to an increased area of dissipation due to
segmentation, which also reduces the flow blockage effect because the air can still flow
through the heatsink channels to some extent even in unfavorable conditions (i.e. flow
perpendicular to the channels). Additionally, each fin acts as a vertex generator, to further
enhance convective heat transfer. Fig. 3c shows a visual depiction of fin geometric
parameters.

Fig. 3.

For convenience, during the testing stage the fins were attached to a flat absorber
plate of the same material using conductive epoxy resin with metal powder (Devcon R2-

11
42). This arrangement was then attached to the PV panel using thermal paste, making it
possible to remove the heatsink from the PV panel for different analyses.
An experimental setup was designed to evaluate the thermal performance of the
heatsinks based on solar irradiance, wind flow direction and PV panel temperature. The
experimental scheme is shown in Fig. 4, and consists mainly of a solar panel, fan, thermal
imager and artificial radiation source (solar simulator).

2.3.2. Experimental setup

The solar simulator was developed with dimensions of 70 cm x 70 cm, a vertical


adjustment of up to 150 cm, and an effective area of 40 cm x 40 cm (Fig. 4). The effective
radiation area is larger than the surface area of the PV test panel, so the level of irradiance
is considered with the latter, as well as its spatial non-uniformity. For the radiation source
4 halogen lamps of 500 W are used. This technology was chosen because it is a proven
low-cost alternative and for ease of set up [43,44].

Fig. 4.

For calibration, a 40 cm x 40 cm grid with a resolution of 5 cm was placed and


measurements were taken with a pyranometer at 81 points. The irradiance level can be
adjusted by varying the distance between the lamps and the PV panel. At 70 cm, an
average irradiance of 819 W/m2 was obtained in the effective area with a non-uniformity
percentage of 5.8% (Fig. 5), which is in the range of commercial solar simulators [45]. A
70 cm distance was determined as adequate in terms of the irradiance/temperature
relation, since reducing the distance to obtain a higher irradiance resulted in a high PV
panel surface temperature (> 90 °C).

Fig. 5.

12
The wind effect was emulated with a fan (Fig. 4). For the best air flow simulation, the
fan was moved to a distance (1.7 m) where the wind speed reached an average of 2 m/s
on the surface of the PV panel (measured with digital anemometer), as the wind speed
tends to decrease when moving away from the fan, both frontally and radially. Speed
measures were taken in 18 points over the surface of the PV module, 9 on the upper
surface, and 9 on the backside. Fig. 4 shows the measure points (red markers) over the
PV module. To evaluate the main wind attack directions (0°, 45° and 90°), a solar
simulator suspension system was attached to the ceiling to facilitate its positioning on a
test table, in addition to giving stability by adjusting the line hangers. This allowed the air
to flow freely and the fan to remain in a fixed position, so it was only necessary to rotate
the simulator to evaluate the different directions of wind attack.

2.3.3. Experimental procedure

Two identical PV panels were used, both 15 W power DSP-15P with polycrystalline
silicon cells and dimensions of 35 cm x 36 cm. One was used as a reference case (no
heatsink), and the other for use with both segmented and conventional fin profile
heatsinks. Initially, base thermal measurements (without heatsink) were made to identify
differences that could be due to manufacturing variations of the PV modules in function
of their mean surface temperature. During several test iterations there were no
differences based on the observation of the thermal image, regarding the average surface
temperature from both modules.
PV modules reach thermal equilibrium in about 15 minutes under the lamp’s radiation;
however, thermal images were taken at minute 20. For the collection of temperature data,
a Fluke Ti-S40 thermographic camera mounted on a tripod was used. When taking
thermographic images, an apparent temperature effect was observed on the surface of
the PV Panel due to the panel reflecting the lamps on the surface. This was also attributed
to the proximity with the incandescent filaments from the simulator's halogen lamps. To
eliminate this effect, a shading screen was added immediately before taking an image (5
seconds). The thermal camera was configured to measure a surface with an emissivity

13
of 0.91 (corresponding to the PV panel), and room temperature was controlled at an
average temperature of 28 °C.

3. Results and discussion

3.1. Simulation results

Principal results regarding thermal performance are presented in Table 4. The


reference PV module, as well as 3 variants of heatsink with discontinuous fin profiles were
evaluated at 5 wind flow directions (0°, 22.5°, 45°, 67.5° and 90°). The performance of
the heatsink varies in function of wind flow direction and therefore, a heatsink can perform
better or worse for a given direction, thus affecting the temperature of the PV module.
The general trend indicates that the temperature is lower when the wind flows diagonally
to the fins (22.5° - 45°), and higher when the flow is perpendicular to them (90°). The
maximum temperature reduction obtained is around 7.28 °C (16%), which corresponds
to an increase in power of 0.5 W (3.66%). Discontinuous profiles of fins 20 mm and 13.3
mm wide perform better; however, fins of 20 mm width present a more consistent
temperature reduction in all wind flow directions.

Table 4
Summary of simulation results (mean PV module cell temperature and power).
Module temp. (°C) Temp. reduction (°C) Module power (W)
Fin width (mm)
Best case Worst case Best case Worst case Best case Worst case
Reference 45.02 45.49 0 0 13.64 13.61
(no heatsink)
320 38.03 39.18 6.99 6.31 14.12 14.05
(conventional)
20 37.83 38.34 7.19 7.15 14.13 14.1
13.3 37.78 39.25 7.24 6.24 14.14 14.04
10 37.74 39.44 7.28 6.05 14.14 14.03

3.2. Thermal effect varying the fin width

Fig. 6 shows the surface temperature distribution of the PV module with and without
fins, using the following as constants: wind direction at 90°, with fins 40 mm long, spaced
14
at 42 mm from one row to the other and a thickness of 0.71 mm, varying only the fin
width (the fins are represented in Fig. 6f). Temperature contours correspond to the cases
of wind flow at 90°, as this is considered the worst scenario in which the heatsinks could
have the greatest pressure losses. The report tool was used during post-processing to
obtain the exact values of maximum and lowest temperature on the displayed surfaces.
In Fig. 6a the thermal distribution of the reference panel without a heatsink, has a
gradient between top and bottom sides of 13.66 °C. In Fig. 6b the gradient using 20 mm
fin width is reduced to 7.7 °C, to 9.2 °C in Fig. 6c (13.3 mm fin width), to 9.7 °C in Fig. 6d
(10 mm fin width), and to 6.4 °C in Fig. 6e (320 mm fin width, straight conventional
heatsink). In qualitative terms, when reducing the width of the fins, the region with the
highest temperature tends to increase (yellow tones). It is also worth mentioning that,
although the conventional heatsink has the lowest thermal gradient (Fig. 6e), it tends to
concentrate heat in the central region of the PV panel, and therefore its average
temperature does not turn out to be the lowest, which is obtained with fins of 20 mm width
for 90° wind flow direction (39.65 °C).
Results show that adding the heatsink contributes to an increase in temperature
homogeneity on the surface of the PV module by reducing the thermal gradient, which
benefits its performance. However, greater fin segmentation (fin width < 20 mm) does not
correlate to higher temperature reduction. This is due to hydraulic characteristics which
is discussed in the next sections. The conventional heatsink tends to trap heat in its
middle section. This is mainly due to the increase in pressure drop caused by the
continuous fin profile, which blocks and/or redirects the air flow out of the heatsink
channels.

Fig. 6.

3.3. Effect of airflow direction

Thermal response, in relation to airflow direction and fin width, is shown in Fig. 7.
Temperature level corresponds to the mean PV cell layer temperature of the model.
Generally, it is observed that when the airflow goes from parallel to diagonal to the fins

15
(0° - 45°), the behavior of the heatsinks is rather similar, with a slight tendency for the
temperature to decrease with greater fin segmentation, i.e., reducing fin width. However,
when the flow goes from diagonal to perpendicular to the fins (45° - 90°), the 20 mm wide
fins show a considerable advantage in temperature reduction. In effect, when considering
the average for all cases of study, the 20 mm fin width profile heatsink performs the best
due to having the most stable thermal response.

Fig. 7.

Parallel airflow (0°) could be considered optimal, as air flows unrestricted and has
good contact with fin surfaces. However, this does not significantly increase flow
turbulence. When the flow angle increases to a certain extent (0° - 45°), the most
favorable cases are presented in which the orientation of the fins allows a less restricted
flow; also, there is reasonable air contact with the dissipation surfaces, and the turbulence
in the flow is intensified. This accounts for the lowest temperatures in the PV system, as
can be seen in Fig. 7. It is evident that with a flow perpendicular (90°) to the fins, the
thermal performance of the heatsink is lower. This is due to the less than optimal
interaction of the fins with the airflow, in which the flow is restricted and losses of pressure
and speed are greater. Further analysis is carried out in the next section regarding
velocity, turbulence and air pressure behavior inside the heatsink channels.

3.4. Hydraulic performance

A discontinuous fin profile has two main advantages over a conventional geometry:
an increased heat dissipation surface, and better adaptation to multidirectional flow.
Pressure drop is a variable of great importance that directly affects the thermal
performance of a heatsink. Abrupt pressure drops can cause low flow uniformity and
adversely affect cooling capacity. Pressure drop is significantly reduced overall using a

16
discontinuous fin profile, as shown in Fig. 8a. In a parallel airflow situation (0°), the
conventional geometry shows a slightly lower pressure drop. However, thermal
performance is better for the segmented profiles, owing to their increased dissipation
surface which also enhances turbulence to a certain extent (Fig. 8b). When the airflow is
diagonal (45°), the conventional heatsink begins to show higher pressure drops, as
continuous parallel fins tend to block the flow and affect its speed. For the most critical
scenario, when the air flows perpendicular to the fins (90°), pressure drop is reduced up
to 49% using a segmented fin profile. This case is further discussed due to being the
worst for all heatsinks regarding hydraulic performance.

The Nusselt number is a dimensionless quantity that indicates the relationship


between convective and conductive heat transfer. In this case, the Nusselt number is
useful to identify in which heatsinks the convective heat transfer is more predominant
and, therefore, which designs are more effective in reducing the temperature of the PV
module. The average Nusselt number is defined as [46]:

𝑞𝐴ℎ𝐷ℎ
𝑁𝑢 = 𝑘𝐴𝑐(𝑇𝑖,𝑎𝑣 ― 𝑇𝑓,𝑎𝑣) (6)

where 𝑞 is the heat flux density, 𝑘 is the thermal conductivity, 𝐴ℎ is the area of heater
surface, 𝐴𝑐 is the interface area of fluid and solid, 𝑇𝑖,𝑎𝑣 is average interface temperature
and 𝑇𝑓,𝑎𝑣 is the average fluid temperature. 𝐷ℎ is the hydraulic diameter:

4𝐴𝑠
𝐷ℎ = 𝑃 (7)

where is 𝐴𝑠 the cross-sectional area of the heatsink channel and 𝑃 is the wetted perimeter
of the channel.

In Fig. 8c can be observed that the Nusselt number has a direct inverse relationship
with the temperature of the PV module (Fig. 8b). The segmented fin profiles present a
higher predominance of convective heat transfer compared to the continuous profile when
the airflow goes from 0º to 45º. This behavior is reflected in the temperature of the PV
module and shows that the segmented profiles are more effective in reducing temperature
in the same airflow range (Fig. 8b). However, as the flow angle increases from 45º to 90º,
the segmented profiles display a decrease in their convective heat transfer capacity

17
(especially the 13.3 mm and 10 mm fin widths). This is attributed to the suboptimal fin
orientation, as it produces flow deflection and separation, reducing the contact with the
dissipation surfaces. In comparison, the continuous fin profile (320 mm) generally shows
less convective capacity, however, its performance is higher at a flow angle of 67.5 ° (Fig.
8c). This is due to the tendency of the continuous fin profile to redirect part of the flow into
the heatsink channels (i.e., the flow that impacts the side of the heatsink), which produces

favorable results for an airflow between ~60° to 78º, despite causing higher pressure

losses. Nevertheless, when the flow is normal to the fins (90º), the segmented profile with
a 20 mm fin width is more effective in reducing the temperature of the PV module, as it
reduces flow stagnation observed in the continuous profile (Fig. 8b). It can also be
observed that the behavior of the 20 mm width fin profile differs from the rest of the
segmented profiles for the 67.5º to 90º airflow range. This indicates that a more stable
thermal response could be obtained by increasing fin width, by combining the advantages
of allowing a less restricted flow when the air flows frontally to the fins, and, to some
extent, redirect the flow towards the heatsink channels when the air hits the side of the
heatsink. It is worth mentioning that the average Nusselt number for the PV panel without
a heatsink ranges from 33 to 34 depending on airflow direction, thus adding a heatsink
considerably improves its convective heat transfer.

Fig. 8.

In Fig. 9, pressure and temperature contour planes corresponding to the working fluid
within each heatsink are compared in function of their fin width at an air flow of 90° (from
top to bottom). Segmented heatsinks have lower and less sudden pressure loss
compared to a conventional geometry (Fig. 9d), due to an improved design which allows
a less restrictive airflow, and causes the downstream fluid to flow faster and retain a
higher heat extraction capacity. This makes the proposed design more thermally stable,
as the conventional design tends to cause stagnation within its channels, heating the
upstream working fluid (Fig. 9d). However, greater segmentation of the fins does not have

18
a considerable effect on pressure drop. In absolute terms, reducing the fin width from 20
mm to 10 mm only decreases the pressure drop around 3.33 Pa, which does not justify
further segmentation of the fins.

Fig. 9.

Fig. 10 shows a velocity streamline in the same plane where pressure/temperature


contours were displayed. Airflow is at 90° (from top to bottom). It is evident that the
conventional heatsink blocks the airflow and redirects it outwards of the heatsink
channels. This causes a low speed zone in its central region which is also related to the
negative pressure zone in the first channel of the heatsink (from top to bottom) as
observed in Fig 9d, as well as with the high temperature zone shown in Fig. 6e.
On the other hand, segmented heatsinks allow a freer airflow and reduce its speed
more gradually. It is noteworthy that a greater presence of eddies is observed (Figs. 10a-
10c), indicating a higher turbulence intensity that favors convective heat transfer by
increasing fluid mixing. The 20 mm width fin heatsink features higher flow velocity in its
rear channels as well as the higher turbulence intensity within segmented profiles,
identifiable with green/light blue tones. This explains its superior thermal performance as
observed in Fig. 8b and lower presence of hot spots in Fig. 9a.

Fig. 10.

3.5. Experimental results

For experimental tests, the heatsink which showed the best results in the theoretical
evaluations was built (fin width and length of 20 mm and 40 mm, respectively), and
compared with its conventional geometry counterpart (constant straight fin). The
temperatures reported in Table 5, both from the CFD model and those obtained
experimentally, correspond to the surface average of the PV panel, given that the thermal

19
imager measures the surface temperature of the module (not the PV cells). Experimental
values correspond to the averages of 4 thermography sessions. An experimental sample
of thermal images with airflow at 90° are shown in Fig. 11.

Fig. 11.

In the case of 90° airflow, a good correspondence with the theoretical model is
observed. The trend in which the segmented heatsink achieves the greatest temperature
reduction was consistently repeated. A greater presence of heat is observed when using
a continuous fin profile (Fig. 11b), which is related to the higher pressure drop in the
middle section of the heatsink and the loss in downstream heat extraction capacity of the
working fluid due to stagnation (see Fig. 9d). It is worth mentioning that the higher
temperature spot on the PV panel corresponds to the junction box (red/white tones).

Table 5
Mean surface temperature of the PV panel.
Model (°C) Experimental (°C)
Airflow No Conventional Segmented No SD Conventional SD Segmented SD
heatsink heatsink
90° 46.28 39.94 39.65 45.84 0.24 41.23 0.24 40.7 0.28
45° 45.8 39.44 39.18 45.7 0.21 41.37 0.27 41.27 0.25
0° 46.15 40.45 39.58 45.84 0.22 41.4 0.26 40.88 0.29

3.6. Comparison of simulation and experiment

In general, it is observed that experimental results are in good agreement with the
model. Additionally, experimental values are quite consistent as the standard deviation
for each case of study is within 0.3 °C (Table 5). However, when comparing numerical
with experimental results, there is a higher deviation for the cases with a heatsink, where
the relative error reaches up to 5.33%. This can be explained in part to the appearance
of hot spots, as can be seen in Fig. 11b, where a group of hot cells can be seen in the
middle part of the panel which also affects the average temperature of the module
surface. The computational model does not replicate this behavior. On the other hand,
the average relative error for the reference cases (without heatsink) is 0.61%. For the

20
critical case (airflow at 90°), the proposed heatsink reduces the temperature by 6.63 °C
and 5.14 °C, numerically and experimentally, respectively. Also, a lower temperature can
be observed towards the middle of the PV module with a segmented heatsink (depicted
by light blue tones). In all cases, the segmented heatsink achieves the greatest
temperature reduction with greater distribution homogeneity, by reducing the intensity of
hot spots (Fig. 11c).

4. Conclusions

This work reports the numerical and experimental results of the evaluation of a
passive cooling system with discontinuous fin profiles for PV modules, optimized for multi-
directional wind flow. The results indicate that the proposed design is more suitable for
scaling compared to a conventional geometry heatsink as pressure drop is reduced
considerably and thermal response is more consistent considering a variable direction
wind flow. This is significant, since when scaling these systems to use them in higher
capacity PV modules, the length of the heatsink channels is increased, i.e., the distance
that the fluid must go through the heatsink increases, incurring pressure losses. Also, the
results obtained show that the proposed system is comparable to the works reported in
the literature, and, in certain cases, achieves a better performance, in terms of PV
operating temperature reduction.
A 20 mm width fin was found to be the optimal, as thinner fins reduced heat transfer
in exchange of a marginal reduction in pressure losses. Based on these results, the study
of wider fins (>20 mm) is justified in future works, as this could further improve the
performance of the heatsink. Regarding hydraulic performance, the proposed heatsink
model showed lower pressure losses compared to a conventional geometry (continuous
fin profile) by reducing the flow velocity losses in the channels, which, in addition to the
design geometry, also increases turbulence promoting convective heat transfer. For the
worst-case scenario (airflow at 90°), pressure loss was reduced up to 49%.
Theoretical temperature reductions are in good agreement with experimental
observations, within the 5-7 °C range, this results in an increase in power yield of up to
2.96% when compared to a PV module with no heatsink. Such a reduction in temperature

21
avoids losses in the electrical conversion efficiency under high irradiance conditions.
Furthermore, this thermal aging reduction could extend the life span of the PV panels,
especially when using solar tracking systems which are exposed to direct solar radiation
for longer periods.
In conclusion, the proposed heatsink considerably reduced the thermal cycle to which
the PV solar panel was exposed in normal conditions, even in cases where the airflow
direction was not optimal. Long-term in situ testing would provide a detailed performance
response to complex natural weather conditions. In addition, the optimization and
evaluation of other heatsink parameters, such as fin spacing and angle, require further
work beyond the scope of this study.

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 work was supported by the project SENER-CONACYT S0019-2014-01, grant
number 254667. We thank Francisco Koh-Dzul for his technical support.

References
[1] S. Armstrong and W. Hurley, A thermal model for photovoltaic panels under
varying atmospheric conditions, Appl. Therm. Eng. 30 (2010) 1488-1495.
[2] D. Philips and W. Warmuth, Fraunhofer ISE: Photovoltaics Report updated: 14
March 2019, no. March (2019).
[3] S. Nižetić, A. Papadopoulos, and E. Giama, Comprehensive analysis and general
economic-environmental evaluation of cooling techniques for photovoltaic panels,
Part I: Passive cooling techniques, Energy Convers. Manage. 149 (2017) 334-354.
[4] H. Bahaidarah, A. Subhan, P. Gandhidasan, and S. Rehman, Performance
evaluation of a PV (photovoltaic) module by back surface water cooling for hot
climatic conditions, Energy 59 (2013) 445-453.
[5] D. Du, J. Darkwa, and G. Kokogiannakis, Thermal management systems for
Photovoltaics (PV) installations: A critical review, Solar Energy 97 (2013) 238-254.
[6] Z. Ling et al., Review on thermal management systems using phase change
materials for electronic components, Li-ion batteries and photovoltaic modules,
Renewable and Sustainable Energy Rev. 31 (2014) 427-438.

22
[7] B. Zhao, W. Chen, J. Hu, Z. Qiu, Y. Qu, and B. Ge, A thermal model for amorphous
silicon photovoltaic integrated in ETFE cushion roofs, Energy Conversion Manage.
100 (2015) 440-448.
[8] M. I. Sohel, Z. Ma, P. Cooper, J. Adams, and R. Scott, A dynamic model for air-
based photovoltaic thermal systems working under real operating conditions, Appl.
Energy 132 (2014) 216-225.
[9] A. Ndiaye, C. M. Kébé, A. Charki, P. A. Ndiaye, V. Sambou, and A. Kobi,
Degradation evaluation of crystalline-silicon photovoltaic modules after a few
operation years in a tropical environment, Sol. Energy 103 (2014) 70-77.
[10] N. Kahoul, M. Houabes, and M. Sadok, Assessing the early degradation of
photovoltaic modules performance in the Saharan region, Energy Conversion
Manage. 82 (2014) 320-326.
[11] R. Edgar, S. Cochard, and Z. Stachurski, A computational fluid dynamic study of
PV cell temperatures in novel platform and standard arrangements, Sol. Energy
144 (2017) 203-214.
[12] J. H. Wohlgemuth, D. W. Cunningham, P. Monus, J. Miller, and A. Nguyen, Long
term reliability of photovoltaic modules, IEEE 4th World Conference on
Photovoltaic Energy Conference 2 (2006) 2050-2053.
[13] E. Radziemska and E. Klugmann, Thermally affected parameters of the current–
voltage characteristics of silicon photocell, Energy Conversion Manage. 43 (2002)
1889-1900.
[14] X. Han, Y. Wang, and L. Zhu, The performance and long-term stability of silicon
concentrator solar cells immersed in dielectric liquids, Energy Conversion Manage.
66 (2013) 189-198.
[15] S. Krauter, Increased electrical yield via water flow over the front of photovoltaic
panels, Sol. Energy Mater. Sol. Cells 82 (2004) 131-137.
[16] H. Teo, P. Lee, and M. N. A. Hawlader, An active cooling system for photovoltaic
modules, Appl. Energy 90 (2012) 309-315.
[17] S. Nižetić, D. Čoko, A. Yadav, and F. Grubišić-Čabo, Water spray cooling
technique applied on a photovoltaic panel: The performance response, Energy
Conversion Manage. 108 (2016) 287-296.
[18] J. C. Mojumder, W. T. Chong, H. C. Ong, and K. Leong, An experimental
investigation on performance analysis of air type photovoltaic thermal collector
system integrated with cooling fins design, Energy Build. 130 (2016) 272-285.
[19] M. C. Browne, B. Norton, and S. J. McCormack, Heat retention of a
photovoltaic/thermal collector with PCM, Sol. Energy 133 (2016) 533-548.
[20] L. Rekha, C. V. Vazhappilly, and C. Melvinraj, Numerical Simulation for Solar
Hybrid Photovoltaic Thermal Air Collector, Procedia Technol. 24 (2016) 513-522.
[21] M. Hasanuzzaman, A. Malek, M. Islam, A. Pandey, and N. Rahim, Global
advancement of cooling technologies for PV systems: A review, Sol. Energy 137
(2016) 25-45.
[22] M. Chandrasekar and T. Senthilkumar, Experimental demonstration of enhanced
solar energy utilization in flat PV (photovoltaic) modules cooled by heat spreaders
in conjunction with cotton wick structures, Energy 90 (2015) 1401-1410.

23
[23] C. G. Popovici, S. V. Hudişteanu, T. D. Mateescu, and N.-C. Cherecheş, Efficiency
Improvement of Photovoltaic Panels by Using Air Cooled Heat Sinks, Energy
Procedia 85 (2016) 425-432.
[24] M. Abd-Elhady, M. Fouad, and T. Khalil, Improving the efficiency of photovoltaic
(PV) panels by oil coating, Energy Conversion Manage. 115 (2016) 1-7.
[25] A. Al Tarabsheh, S. Voutetakisb, A. Ι. Papadopoulosb, P. Seferlisb, I. Etiera, and
O. Saraereha, Investigation of Temperature Effects in Efficiency Improvement of
Non-Uniformly Cooled Photovoltaic Cells, Chem. Eng. Trans. 35 (2013).
[26] A. Elbreki, M. Alghoul, K. Sopian, and T. Hussein, Towards adopting passive heat
dissipation approaches for temperature regulation of PV module as a sustainable
solution, Renewable Sustainable Energy Rev. 69 (2017) 961-1017.
[27] A. El Mays et al., Improving photovoltaic panel using finned plate of aluminum,
Energy Procedia 119 (2017) 812-817.
[28] F. Bayrak, H. F. Oztop, and F. Selimefendigil, Effects of different fin parameters
on temperature and efficiency for cooling of photovoltaic panels under natural
convection, Sol. Energy 188 (2019) 484-494.
[29] F. Grubišić-Čabo, S. Nižetić, D. Čoko, I. M. Kragić, and A. Papadopoulos,
Experimental investigation of the passive cooled free-standing photovoltaic panel
with fixed aluminum fins on the backside surface, J. Cleaner Prod. 176 (2018) 119-
129.
[30] A. Abdel-raheimAmr, A. Hassan, M. Abdel-Salam, and A. M. El-Sayed,
Enhancement of photovoltaic system performance via passive cooling: Theory
versus experiment, Renewable Energy 140 (2019) 88-103.
[31] F. Selimefendigil, F. Bayrak, and H. F. Oztop, Experimental analysis and dynamic
modeling of a photovoltaic module with porous fins, Renewable Energy 125 (2018)
193-205.
[32] F. G. Čabo, S. Nižetić, E. Giama, and A. Papadopoulos, Techno-economic and
environmental evaluation of passive cooled photovoltaic systems in Mediterranean
climate conditions, Appl. Therm. Eng. 169 (2020) 114947.
[33] Alpha Novatech Inc., Online Catalog - LPD Series.
https://www.alphanovatech.com/en/cat_lpde.html, 2008 (accessed 30 June 2020).
[34] Digi-Key Electronics, Thermal - Heat Sinks.
https://www.digikey.com/products/en/fans-thermal-management/thermal-heat-
sinks/219, 2020 (accessed 30 June 2020).
[35] HeatsinkUSA, 10.000" Wide Extruded Aluminum Heatsink.
https://www.heatsinkusa.com/10-000-wide-extruded-aluminum-heatsink/, 2020
(accessed 30 June 2020).
[36] Mouser Electronics, Heat Sinks. https://www.mouser.com/Thermal-
Management/Heat-Sinks/_/N-5gg0, 2020 (accessed 30 June 2020).
[37] T-Global Technology Co., Heat Sink (>30mm).
https://www.tglobalcorp.com/heat-sink-3 (accessed 1 July 2020).
[38] Digi-Key Electronics, 122256 Wakefield-Vette.
https://www.digikey.com/product-detail/en/advanced-thermal-solutions-inc/ATS-
EXL6-254-R0/ATS2204-ND/5848412, 2020 (accessed 1 July 2020).

24
[39] Digi-Key Electronics, ATS-56006-C3-R0. https://www.digikey.com/product-
detail/en/advanced-thermal-solutions-inc/ATS-56006-C3-R0/ATS1337-
ND/1285051, 2020 (accessed 1 July 2020).
[40] J. Hernandez-Perez, J. Carrillo, A. Bassam, M. Flota-Banuelos, and L. Patino-
Lopez, A new passive PV heatsink design to reduce efficiency losses: A
computational and experimental evaluation, Renewable Energy 147 (2020) 1209-
1220.
[41] ANSYS Inc., ANSYS fluent theory guide 15.0, ANSYS, Canonsburg, PA, 2013.
[42] K. Kant, A. Shukla, A. Sharma, and P. H. Biwole, Thermal response of poly-
crystalline silicon photovoltaic panels: Numerical simulation and experimental
study, Sol. Energy 134 (2016) 147-155.
[43] Y. Irwan, W. Leow, M. Irwanto, A. Amelia, N. Gomesh, and I. Safwati, Indoor test
performance of pv panel through water cooling method, Energy Procedia 79 (2015)
604-611.
[44] H. Moria, T. Mohamad, and F. Aldawi, Radiation distribution uniformization by
optimized halogen lamps arrangement for a solar simulator, J. Sci. Eng. Res. 3
(2016) 29-34.
[45] Standard Specification for Solar Simulation for Photovoltaic Testing, ASTM E927-
10, A. International, West Conshohocken, PA, 2015.
[46] D. Ma, G. Xia, Y. Li, Y. Jia, and J. Wang, Effects of structural parameters on fluid
flow and heat transfer characteristics in microchannel with offset zigzag grooves
in sidewall, Int. J. Heat Mass Transfer, 101 (2016) 427-435.
Figure captions

Fig. 1. Common heatsink designs in the industry, (a) Vertical fins [38], (b) Angled fins
[39].
Fig. 2. (a) Computational domain, (b) Model zones, (c) Mesh cross-section detail.
Fig. 3. PV cooling system, (a) Conventional heatsink, (b) Proposed heatsink, (c) Fin
parameters.
Fig. 4. Work methodology, (I) Computational methodology, (II) Numerical data analysis,
(III) Experimental methodology.
Fig. 5. Irradiance distribution on the effective area of the solar simulator.
Fig. 6. Simulation of temperature distribution of the PV module for different fin widths, (a)
No heatsink, (b) 20 mm, (c) 13.33 mm, (d) 10 mm, (e) 320 mm, (f) Fin orientation and
airflow direction.
Fig. 7. Modeled response of PV panel temperature by varying airflow direction and fin
number.
Fig. 8. Fin width comparison for airflow at 0°, 45° and 90°, (a) Heatsink pressure drop,
(b) PV module mean temperature, (c) Average Nusselt number of the fluid boundary with
the heatsink surface.
Fig. 9. Pressure/temperature distribution of the working fluid within the heatsinks using
different fin widths, (a) 20 mm, (b) 13.33 mm, (c) 10 mm, (d) 320 mm. Airflow attack at
90° from top to bottom.
25
Fig. 10. Velocity field of the heatsinks using different fin widths, (a) 20 mm, (b) 13.33 mm,
(c) 10 mm, (d) 320 mm.
Fig. 11. Surface temperature distribution of the PV panel, (a) No heatsink, (b)
Conventional heatsink, (c) Segmented fin heatsink, (d) Airflow direction.

26
Highlights

 PV passive cooling system is improved by modifying heatsink geometric


parameters.
 A discontinuous finned heatsink design is used for multi-directional airflow.
 PV module temperature reduced up to 5.5 °C in experimental testing.
 The proposed heatsink design reduced pressure drop by up to 49%.
 A discontinuous fin heatsink profile resulted in a more stable thermal response.

27
Declaration of interests

☒ 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.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

None

28

You might also like