You are on page 1of 11

728

ARTICLE
Numerical simulation of wind loading on ground-mounted
solar panels at different flow configurations
M. Shademan, R.M. Barron, R. Balachandar, and H. Hangan
Can. J. Civ. Eng. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF CONNECTICUT on 10/11/14

Abstract: Three-dimensional Reynolds-Averaged Navier-Stokes simulations have been carried out to evaluate the flow past
ground-mounted solar panels at different flow configurations. Initially, the flow past a stand-alone solar panel consisting of four
individual sub-panels in a 2×2 arrangement is considered. The effects of the lateral gap spacing between sub-panels, the ground
clearance, and the wind direction on the wind loading of the full panel have been analyzed. Simulations of the flow past solar
panels in an arrayed configuration are also conducted to investigate the effect of longitudinal spacing between the panels on the
wind loading. Results from the flow past the stand-alone panel reveal that the structure experiences maximum values of wind
loading at two azimuthal wind directions of ␪ = 0° and 180°. The results also show that the two bottom panels experience larger
mean wind loading compared to the top panels. The introduction of gap spacing between the panels changes the flow structure
in the wake region, contributing to the formation and shedding of additional vortices, which increase in size with increased gap
spacing. Although the introduction of gap spacing results in reduced mean wind forces, it produces regions that experience large
wind loading. Increase of the ground clearance also causes larger mean wind loading on the panels. A proper choice of
longitudinal spacing between the panels forming the array can significantly reduce the lift by taking advantage of the sheltering
effect.

Key words: CFD, solar panel, wind loading, gap spacing, ground clearance, array.
For personal use only.

Résumé : Des simulations tridimensionnelles d’équation de Navier-Stokes à moyenne de Reynolds ont été réalisées afin
d’évaluer le mouvement de l’air passant près de panneaux solaires installés sur le sol pour différentes configurations
d’écoulement. Au départ, le débit d’air circulant près d’un panneau solaire autonome comportant quatre sous-panneaux
individuels selon une configuration de 2 sur 2 a été considéré. Les effets de l’espacement latéral entre les sous-panneaux, la
distance au sol et la direction des vents sur la charge exercée par le vent sur le panneau complet ont été analysés. Des simulations
des débits d’air passant autour des panneaux solaires dans une configuration en réseau ont également été réalisées pour
examiner l’effet de l’espacement longitudinal entre les panneaux sur la charge exercée par le vent. Les résultats pour le débit
d’air circulant près du panneau solaire autonome montrent que la structure subit des valeurs maximales de charge exercée par
le vent pour deux directions de vent azimutales de ␪ = 0° and 180°. Les résultats montrent aussi que les deux panneaux du bas
subissent une plus forte charge exercée par le vent par rapport aux panneaux du haut. L’introduction d’un espacement entre les
panneaux change la structure du débit d’air dans le sillage, contribuant à la formation et à la perte de vortex additionnels dont
les dimensions augmentent en fonction de l’espacement. Bien que l’introduction d’un espacement réduise les forces moyennes
de vent, elle produit des régions qui connaissent de fortes charges exercée par le vent. L’augmentation de la distance du sol cause
aussi une charge exercée par le vent supérieure sur les panneaux. Un bon choix de l’espacement longitudinal entre les panneaux
formant le réseau peut réduire significativement le soulèvement exercé en prenant avantage de l’effet d’abri de protection.
[Traduit par la Rédaction]

Mots-clés : mécanique des fluides numérique, panneau solaire, charge exercée par le vent, espacement, distance au sol, réseau.

1. Introduction Therefore, a good estimation of the structural resistance to wind


effects and the reduction of aerodynamic loads on solar panels is
The use of solar collectors to deliver an economically feasible
essential for preventing potential damage. To improve the design,
green energy resource has increased in recent years. To facilitate
engineers need to enhance their understanding of the aerody-
ease of maintenance and enhance air ventilation, solar panels are namics of these systems.
generally mounted with a ground clearance greater than 0.5 m. Several studies have been carried out to investigate the role of
The lateral interspace (gap) between the panels also varies. Given various parameters on the wind loading of solar panels. Wind
the large surface area of the panels, extreme wind loads acting direction, inclination angle of the panels, ground clearance, and
on the panels may cause serious structural and (or) mechanical fail- sheltering are some of the parameters that have been discussed by
ure. Field experience has shown that breakage of solar panels is researchers. The main objectives in these studies were to evaluate
mainly due to high wind loading and the associated deflections. the critical wind loading conditions and to minimize their effects.

Received 28 November 2013. Accepted 22 June 2014.


M. Shademan. Department of Mechanical, Automotive and Materials Engineering, University of Windsor, Windsor, ON N9B 3P4, Canada.
R.M. Barron. Department of Mechanical, Automotive and Materials Engineering, University of Windsor, Windsor, ON N9B 3P4, Canada; Department of
Mathematics and Statistics, University of Windsor, Windsor, ON N9B 3P4, Canada.
R. Balachandar. Department of Mechanical, Automotive and Materials Engineering, University of Windsor, Windsor, ON N9B 3P4, Canada;
Department of Civil and Environmental Engineering, University of Windsor, Windsor, ON N9B 3P4, Canada.
H. Hangan. WindEEE Research Institute, Western University, London, ON N6A 5B9, Canada.
Corresponding author: Ram Balachandar (e-mail: rambala@uwindsor.ca).

Can. J. Civ. Eng. 41: 728–738 (2014) dx.doi.org/10.1139/cjce-2013-0537 Published at www.nrcresearchpress.com/cjce on 25 June 2014.
Shademan et al. 729

To evaluate the effect of wind direction, Mohapatra (2011) car- Fig. 1. (a) Geometry of a set of stand-alone solar panel in a 2×2
ried out wind tunnel tests on ground-mounted photovoltaic track- arrangement, (b) azimuthal angle (␪), (c) inclination angle (A), and
ers. He noticed that the largest wind loading occurred when the (d) arrayed panels.
wind flow was perpendicular to the panels. Shademan and Hangan
(2009, 2010) carried out computational fluid dynamics (CFD) anal-
yses to determine the wind loads on a set of solar panels consist-
ing of 12 individual panels in a 3×4 arrangement. They performed
simulations for different azimuthal and inclination angles and
found that the entire structure experienced the maximum force
when the wind flow was perpendicular to the panels, either from
Can. J. Civ. Eng. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF CONNECTICUT on 10/11/14

the front or from the back face. In addition, they observed that the
corner panels were the more critical ones in terms of aerodynamic
forces exerted on the structure. These simulations have been ex-
tended by Jubayer and Hangan (2012) and further benchmarked with
experimental results by Jubayer et al. (2012).
Kopp et al. (2002), using wind tunnel testing, found that the
highest wind loads were due to the vortex shedding from in-line
panels and observed that the peak system moment occurred at in-
coming wind angles near the diagonals of the panels. Bitsuamlak
et al. (2010) carried out studies on ground-mounted solar panels. They
noticed that the maximum wind loading occurs at an azimuthal
wind direction of 180°. For one array spacing (1 m), their study dem-
onstrated that the wind loads on the downstream panels are reduced
due to a prominent sheltering effect.
There is little information in the published literature on the
effect that gap spacing between individual panels has on the wind
loading. Wu et al. (2010) found only a very slight increase in load-
ing on heliostat panels for which the gap-to-panel ratio is very
small. In the current study it is of interest to investigate the role of
For personal use only.

gap spacing on the flow structure in the wake of typical solar panels
with larger gap-to-panel ratios, as well as the effect of ground clear-
ance. In addition to this, the effect of different longitudinal spac-
ing between sets of solar panels in an arrayed configuration has
been analyzed to determine an optimal distance at which panels
experience minimum wind loading. To achieve this, a reliable
numerical model is developed to assess the mean aerodynamic
forces acting on a generic set of flat stand-alone solar panels with
different gap spacing and ground clearance, and on solar panels in
an arrayed configuration. Considering that the wind flow is not
limited to only one direction, different azimuthal angles are mod-
eled to determine the critical wind directions for a specific panel-
to-ground inclination angle. As previously noted by Shademan
and Hangan (2009, 2010), a change in the inclination angle does
not significantly influence the locations on the solar panel set
where the wind loading is maximum. Therefore, only one inclina-
tion angle is considered in the current study. At the critical wind
directions, three different gap spacings between individual panels
are then modeled to investigate the effect of this parameter on the
wind loading. For the analysis of the ground clearance effect on
the wind loading, three values for the ground clearance have been
modeled. For the arrayed cases, four rows of solar panels are mod-
eled with different longitudinal spacing between rows.
The general structure of the paper is as follows. Geometry mod-
mounted solar panel set. Several configurations of a 2×2 set of
eling is presented in Section 2. Section 3, which is the validation
generic solar panels have been modeled, with different gap spac-
part of the work, includes comparison of the predicted wind
ing and ground clearance. As shown in Figs. 1b and 1c, the config-
forces with available experimental data. In Section 4, the effect of
urations correspond to one inclination angle (A = 135°) and seven
different parameters such as gap spacing and ground clearance
azimuthal (wind direction) angles (␪ = 0°, 30°, 60°, 90°, 120°, 150°,
have been investigated. The sheltering effect of one set of solar
180°). Various gap spacing (␦ = 0.0, 0.1, 0.2 m) have been considered
panels on other subsequent panels is analyzed in Section 5. at wind direction of ␪ = 0°. The effect of ground clearance on the
wind loading has been analyzed at three different heights (Hp =
2. Numerical method
0.5, 1.5, and 2.5 m). Table 1 presents a summary of the test cases
2.1. Geometry, boundary conditions, and mesh analyzed in the current study.
requirements A three-dimensional rectangular domain shown in Fig. 2a is
The geometry and nomenclature for the set of stand-alone solar established for the computation of the turbulent wind flow past
panels used in this study are illustrated in Fig. 1a. Each individual the solar panels. The recommendations of Franke et al. (2007)
panel is 2.1 m in length, 1.6 m in width, and 0.05 m in thickness, have been used to define the domain dimensions. To accurately
which is representative of the dimensions of a typical ground- model the inlet and outlet, the length of the computational do-

Published by NRC Research Press


730 Can. J. Civ. Eng. Vol. 41, 2014

Table 1. Test cases modeled.


Single set panel
Azimuthal Inclination Gap Ground Arrayed panels; longitudinal
angles (␪) angles (␾) spacings (␦) [m] clearances (Hp) [m] Parameters spacing (S*/⌬)
0° 135° 0.0, 0.1, 0.2 0.5, 1.5, 2.5 at ␦ = 0.2 FD, FL, Mz S*/⌬ = 1, 2, 3
30° 135° 0.0 1.5 FD, FL, Mz N/A
60° 135° 0.0 1.5 FD, FL, Mz N/A
90° 135° 0.0 1.5 FD, FL, Mz N/A
120° 135° 0.0 1.5 FD, FL, Mz N/A
Can. J. Civ. Eng. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF CONNECTICUT on 10/11/14

150° 135° 0.0 1.5 FD, FL, Mz N/A


180° 135° 0.0 1.5 FD, FL, Mz N/A
Note: N/A, not applicable.

Fig. 2. (a) Domain dimensions and boundary conditions, (b) block with hybrid mesh around the panels, and (c) cross section of the mesh.
For personal use only.

main is set at 100 m. A precise wind velocity profile (see eq. (1) grid spacing with h+ < 1 is required to have at least a few cells
below), representative of an atmospheric boundary layer, was im- inside the viscous sub-layer, where h+ = u␶h/v is a non-dimensional
posed at the inlet of the domain. To ensure that the top and side wall distance, h is the normal distance from the wall, v is the
walls have negligible influence on the pressure and velocity fields, kinematic viscosity, u␶ = (␶w/␳)0.5 is the friction velocity, ␶w is wall
the domain extends 46 m in width and 20 m in height. The max- shear stress, and ␳ is the density. Grid independence tests were
imum blockage ratio based on these dimensions is approximately performed with the number of cells being increased in 20% incre-
1.1%, which is significantly lower than the 3% recommended by ments until no noticeable variation in the drag force exerted on
Franke et al. (2007). the panels was observed. The typical number of cells for the cur-
A hybrid grid has been generated using GAMBIT 2.2.30. The rent simulations, after conducting the grid independence tests,
entire solar panel set is imbedded in a rectangular block. An un- was approximately 3.0 × 106. To simulate the wind loading over
structured mesh is produced around the panels inside this rect- solar panels installed in an open terrain, the boundary conditions
angular block, except for the region close to the panels (Fig. 2b). illustrated in Fig. 2a are adopted.
The mesh system is comprised of structured hexahedral elements A wind velocity of 90 km/h at the height of 10 m, which is
around the panels to provide a fine mesh with less skewness near prevalent for most regions in the north-eastern part of the North
the walls. This rectangular block and its mesh are fixed to the American continent, was chosen from the ASCE 7-05 code (2005).
panels and are rotated for each wind direction. A high-density The incoming wind, imposed at the inlet of the computational
mesh is used to capture the high shear stresses generated over the domain, is simulated using the power-law boundary layer profile,
solar panels and also over the ground surface. For the rest of the i.e.
domain, where the wall effect is smaller, a coarser mesh is used.

冉冊
The outer region, exterior to the block, is meshed separately for u(y) y ␣
each wind direction, with fully structured hexahedral elements (1) ⫽
Ug yg
(Fig. 2c).
Different methods are available to account for the effect of a
wall in CFD simulations. In the current simulations, near-wall where Ug is the geostrophic wind velocity, yg is the height of the
regions are modeled using the Low Reynolds Number Modeling atmospheric boundary layer, ␣ is an exponent which is dependent
(LRNM) method which requires a zero roughness height. Surface on the terrain, and y is the distance from the ground (for an open

Published by NRC Research Press


Shademan et al. 731

terrain ␣ = 0.16 and yg = 300 m). This normalized velocity profile Fig. 3. (a) Inlet stream-wise turbulence intensity (Iu) and velocity
(U/Ug) for an open terrain is shown in Fig. 3a along with the exper- (U/Ug), and (b) Inlet, approaching, and incident normalized wind
imental data of ESDU 82026 (ESDU 1982). It should be remarked velocity profiles (U/Up).
that ESDU 82026 provides methods for estimating the variation of
strong hourly-mean wind speeds in the atmospheric boundary
layer. The suggested distribution is particularly applicable in the
calculation of wind loads on buildings and structures. To preserve
the homogeneity of the wind flow approaching the panels, the
upstream length of the domain is chosen to be 20 m, which is the
minimum acceptable value according to the recommendations of
Can. J. Civ. Eng. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF CONNECTICUT on 10/11/14

Blocken et al. (2007). Different velocity profiles including those at


the inlet, approaching (10 m from the panel), and incident (2 m
from the panel) locations are plotted in Fig. 3b. In this figure, Hp is
the solar panel height and Up is the wind velocity at this level. The
figure demonstrates that the change in the upstream profile is
insignificant in the present situation.
The arrayed configuration is shown in Fig. 1d. Four rows of solar
panels with different longitudinal spacing between them (S*/⌬ = 1,
2, 3) are considered for simulation (⌬ is the projection of the chord
of the solar panel on the x-axis and S* is the longitudinal spacing
between two rows). Specifically, we model a 4×1 tandem arrange-
ment and investigate the effect of distance between rows of pan-
els. Assumptions similar to those for the stand-alone solar panels,
including boundary conditions and wall modeling, are used for
the array simulation. However, due to the presence of 4 rows of
panels, the downstream length was extended to 130 m, which
gives the entire domain a total length of 150 m. The total number
of the cells required to achieve grid independent results is
5.0 × 106.
For personal use only.

2.2. Governing equations


Steady Reynolds Averaged Navier-Stokes (RANS) equations are
solved for the velocity and pressure fields for a three-dimensional
incompressible flow. To model the Reynolds stresses resulting
from the Reynolds averaging, the Boussinesq hypothesis is com-
monly employed to relate the Reynolds stresses to the mean ve-
locity gradients. The disadvantage of this method is that in some
turbulence models ␮t is considered as an isotropic scalar quantity,
which significantly influences the results. The Realizable k−␧
model (Shih et al. 1995) and the Shear Stress Transport k−␧ SST
model (Menter 1994) have been used in the current simulations.
For the Realizable k−␧ model, the zonal LRNM was implemented,
with the Realizable k−␧ model in the turbulent region and the where Ret = (␳k/␮␻), ␣ⴱ0 ⫽ 0.024, Rk = 6.0, and a1 = 0.31.
Wolfshtein model (1969) in the viscosity-affected near-wall region. The turbulent kinetic energy and dissipation rate applied at the
The turbulent viscosity is determined from inlet for different turbulence models is based on the equations
presented by Blocken et al. (2007) as
␳C␮k2
(2) ␮t ⫽ uⴱABL
2
␧ (5) k(y) ⫽
兹C␮
Here, k is the turbulent kinetic energy and ␧ is the dissipation rate.
The coefficient C␮ is not a constant value and is dependent on the uⴱABL
3

flow characteristics. The k−␻ SST turbulence model takes into (6) ␧(y) ⫽
␬y
account the low-Re effects in the flow. In this model, for the near-
wall region, a k−␻ model is used and is combined with a standard ⴱ
where uABL is the atmospheric boundary layer friction velocity, ␬ is
k−␧ model in the turbulent zone (Menter (1994)). In the near-wall
the von Karman constant (≈0.40 − 0.42), and C␮ is a model constant
region, the turbulent viscosity is determined from
for the k−␧ model. In the case of the k−␻ SST turbulence model the

共 ␳␻k 兲再 max[(1/␣ ),1 (SF /␣ ␻)] 冎


profiles of k and ␧ are converted into specific dissipation rate ␻ (␻ =
(3) ␮t ⫽ ⴱ
␧/C␮k). The turbulent kinetic energy profile in terms of turbulence
2 1 intensity (Iu) shown in Fig. 3a compares well with the values obtained
from ESDU 83045 (1983), which provides a method for estimating the
Here, F2 is a blending function (Menter (1994)). The parameter ␣ⴱ variation of the expected maximum gust speed with height above
acts as a low-Re correction factor and is determined from the ground.
The finite volume method is used to discretize the governing

冋 册
equations. The QUICK scheme is used for discretizing the convec-
␣ⴱ0 ⫹ (Ret /Rk)
(4) ␣ⴱ ⫽ ␣∞ⴱ tive terms and the STANDARD scheme is used for the pressure
1 ⫹ (Ret /Rk) interpolation. The SIMPLE algorithm developed by Patankar and

Published by NRC Research Press


732 Can. J. Civ. Eng. Vol. 41, 2014

Spalding (1972) is used for the pressure–velocity coupling. The Fig. 4. Comparison of CFD and experiments for (a) Cp obtained from
Reynolds number based on the width of the panel and wind ve- different turbulence models at A = 90°, (b) CD versus wind direction,
locity at the panels’ height is approximately 4×106. FLUENT 6.3.26 (c) CL versus wind direction, and (d) CM versus wind direction.
(2006) is used to solve the governing equations. The simulations
were considered to have converged when no significant change in
the drag force exerted on the solar panels was observed. In all
cases, the residuals for all equations were less than 10−6 at conver-
gence. Seven cases were simulated for the wind direction analysis,
three simulations for ground clearance effect, and three simula-
tions for gap spacing variation.
Can. J. Civ. Eng. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF CONNECTICUT on 10/11/14

3. Validation of numerical model


3.1. Flow over an inclined flat plate
There is a strong similarity between the flow over flat solar
panels and the flow past flat plates. Based on the Reynolds number
in the current simulations, the closest matching benchmark con-
ditions are from the experimental study of Fage and Johansen
(1927). Consequently, a separate simulation matching the geome-
try and boundary conditions of Fage and Johansen (1927) is used to
validate the procedures in the current study. Two-dimensional
RANS simulations were carried out for two inclination angles
(A = 70° and 90°) at ␪ = 0° with the two different turbulence models
(Realizable k−␧ and k−␻ SST). The pressure coefficient CP ⫽
共P ⫺ P∞兲/共0.5␳U∞2 兲 determined from the current simulations (A = 90°)
is compared with the experimental data of Fage and Johansen (1927)
in Fig. 4a (c = W + ␦ is the chord length of the panel). Figure 4a shows
that for this specific geometry the results obtained from the k−␻ SST
model are in better agreement with the experimental results than
For personal use only.

the Realizable k−␧ model. Therefore, k−␻ SST was selected as the
working turbulence model for all subsequent simulations in this
study.

3.2. Wind direction effect


To provide further validation and to determine the wind direc-
tion under which the entire structure of the solar panel set expe-
riences the maximum aerodynamic load, a set of solar panels with
no gap spacing was first considered. The simulations were per-
formed for different wind directions of ␪ = 0°, 30°, 60°, 90°, 120°,
150°, and 180°. From these simulations, the aerodynamic drag (CD),
lift (CL), and moment (CM) coefficients were calculated. These co-
efficients are defined as

FD
(7) CD ⫽
0.5␳UH2A

FL
(8) CL ⫽
0.5␳UH2A

Mz
(9) CM ⫽
0.5␳UH2AW Some discrepancies are noted at oblique wind directions (between
30° and 60°), which may be attributed to the effects of corner
vortices that may not be fully captured by the RANS modeling.
where UH is the wind velocity averaged over the solar panel area From these figures it is clear that ␪ = 0°and ␪ = 180° are the angles
(A). Further, A is the solid area of the panel without gap spacing, FD at which the wind exerts the maximum aerodynamic force on the
and FL are the drag and lift components of the force acting on the entire set of solar panels. These results are consistent with those
panels, respectively, and Mz is the moment of aerodynamic force of Shademan and Hangan (2009, 2010).
acting at the middle of the panel in the z direction. Note that the Figure 4b shows that there is a significant reduction in drag
computed values of CD and CL were determined based on the force as ␪ increases from 0° to 90°. However, the results shown in
values of FD and FL obtained from the calculated pressure distri- Figs. 4c, 4d indicate that there is no significant difference in the
bution on the panel. magnitude of the lift and moment coefficients for 0° < ␪ < 45° and
The experiments carried out by Mohapatra (2011) on flat solar 135° < ␪ < 180°. Nevertheless, the lift and moment coefficients are
panels without gap spacing but bearing some similarity to the large in these ranges. It is well known that most of the structural
present flow configuration and ground clearance, have also been failures associated with solar panels are due to the lift and mo-
used to validate the simulations. Figures 4b, 4c, 4d compare the ment experienced at high wind velocities. The present results
force and moment coefficients, demonstrating that the present confirm that the dominant component of force on the panels is
numerical results compare well with those of the experiments. the lift and moment over a wide range of azimuthal angles, i.e.,

Published by NRC Research Press


Shademan et al. 733

wind directions. Hence, care should be taken to consider the lift nearly unchanged. At a gap spacing of 0.1 m, the upper portion of
and moment when designing solar panel configurations. the lower leeward panel experiences lower pressure (point A)
compared to a gap spacing of 0.2 m, where the lower portion of
4. Results for stand-alone panels the panel (point B) experiences a lower pressure. This flipped
4.1. Gap spacing effect behaviour of the pressure coefficient curves on the leeward face of
Depending on the particular application, solar panels are gen- the lower panel is related to the flow structure, which is discussed
erally installed with different gap spacing between the panels. To in the next section.
study the effect of gap spacing on the aerodynamic forces exerted The pressure differential between the windward and leeward
on the solar panel, 3D RANS simulations were carried out with gap sides is drastically diminished in the gap region only. On evaluat-
spacing of ␦ = 0.0, 0.1, 0.2 m. The solid area of the entire set is kept ing the drag and lift forces on the entire set of solar panels, one
Can. J. Civ. Eng. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF CONNECTICUT on 10/11/14

constant and only the gap spacing is changed. The simulation observes that increasing the gap spacing reduces the mean aero-
assumptions (boundary conditions, mesh, solution scheme, and dynamic forces. This reduction is higher at ␪ = 180° than at ␪ = 0°.
wall treatment) are similar to those stated in the previous sec- This reduction in aerodynamic forces (5%) due to the increase in
tions. For these simulations, the mesh inside the gap spacing was gap spacing is attributed to the change in the flow structure in the
refined to meet the LRNM criterion (h+ < 1), to correctly capture wake of the panels.
the high shear flow through the openings. Since ␪ = 0° and ␪ = 180°
4.1.3. Gap spacing effect on flow structure
are the critical angles in terms of wind loading, these two wind
To get a better understanding of the fluid flow around the pan-
directions were selected for further analysis.
els, snapshots of the mean sectional streamlines for different gap
In subsection 4.1.1, the location on each individual panel where
the wind loading is highest is determined. The effects of introduc- spacing configurations are plotted in Fig. 7. Figures 7a, 7b, 7c show
ing different gap spacing on the pressure distribution and flow the streamlines superimposed with the vorticity contours on the
structure on mid-section planes are analyzed in subsections 4.1.2 x-y plane through the mid-span where the pressure coefficients in
and 4.1.3, respectively. Finally, Section 4.2 presents the effect of Fig. 6 are plotted. Figures 7d, 7e, 7f show the sectional streamlines
different ground clearance on the aerodynamic forces exerted on of the flow on the x-z plane (parallel to the ground) in the middle
the entire structure. of the panels in the upper row, superimposed with the vorticity
contours. As expected, the gap spacing has little effect on the
4.1.1. Wind loading and pressure distribution on individual mean sectional streamlines upstream of the body. The mean stag-
panels nation point occurs close to the lower edge of the panels in all
Since the pressure distributions on the windward and leeward cases. However, with the gap, flow is induced through the gap
For personal use only.

surfaces determine the total aerodynamic force (F ⫽ 冖PdA) on the causing significant changes in the wake. Due to the transport of
panels, the location on the panels where the maximum pressure fluid through the openings, vortices are generated in the near
difference occurs is important. This information can guide engi- wake immediately following the gap. As can be seen in Figs. 7b, 7c,
neers to put more emphasis on the design of panel sections where 7e, 7f, increasing the size of the gap results in expansion of these
the wind loading is expected to be larger. The mean static pres- vortices. Another significant influence of the gap can be observed
sure distribution on the windward and leeward faces of the panels in the size of the dominant vortex behind the panels, which is
with gap spacing of 0.2 m are presented for ␪ = 0° (Figs. 5a, 5b) and formed due to the fluid passing from the top and bottom of the
␪ = 180° (Figs. 5c, 5d), respectively. Note that the numbering of the structure. The dominant vortex decreases in size in both x-y and
panels is in accordance with the format presented in Fig. 1. One x-z planes as the gap spacing increases. Reduction in size of the
can note that there is a monotonic decrease in pressure on the dominant vortex produces a smaller low-pressure region behind
windward face starting from the lower part of each individual the panels, causing smaller aerodynamic forces.
panel towards the upper edge at ␪ = 0° (Fig. 5a) and in the opposite By considering the streamline pattern in the x-y plane for dif-
direction for ␪ = 180° (Fig. 5c). These contours show that at ␪ = 0° ferent gap spacing (Figs. 7a, 7b, 7c) it can be observed that the
the two bottom panels (3 and 4) experience a larger pressure location of the saddle point (S) changes by increasing the gap
difference between windward and leeward surfaces compared to spacing. The saddle point occurs at the location where a parcel of
the top panels (1 and 2). The trend is opposite for ␪ = 180°, where fluid is separated by another part of the flow with opposite veloc-
the top panels are subjected to larger pressure difference. Com- ity gradient causing a large-scale structure in the wake of the bluff
plete and detailed contours of pressure distribution like those body. The location of the saddle point for panels without gap
shown in Fig. 5 are difficult to obtain in an experimental study spacing is at the right corner of the dominant vortex behind the
because, at a typical wind tunnel scale of 1/20 to 1/50, the physical panel (Fig. 7a). By increasing the gap spacing size to 0.1 m in Fig. 7b,
size of the scaled solar panel model cannot host a large number of the size of the dominant vortex decreases and the saddle point
pressure probes. CFD based calculations are therefore useful in moves closer to the panel. In this case, due to the flow through the
producing mean surface pressures, which can lead to better de- gap and generation of the different velocity gradient on the lower
sign guidelines. side of the panel, a second saddle point appears in this region.
Furthermore, at a larger gap spacing (0.2 m), the top saddle point
4.1.2. Gap spacing effect on pressure distribution which was observed in 0.0 and 0.1 m gap cases (Figs. 7a, 7b) disap-
Figure 6 illustrates the pressure distribution on the mid-span of pears and only the one located close to the lower part of the panel
the solar panels (either left or right column, see Fig. 1a) for ␪ = 0° is observed. The movement of the saddle point towards the panel
with different gap spacing. It is important to note that this Cp (Figs. 7a, 7b) is due to the reduction in velocity gradient (vorticity)
distribution is not representative of the pressures over the entire between the flow moving from bottom and top sides, creating a
panels. Comparison of the three curves in Fig. 6 shows that intro- smaller dominant vortex behind the panels.
ducing the gap spacing dramatically influences the pressure dis- As Figs. 7b, 7c, 7e, 7f demonstrate, a portion of the flow passing
tribution on both the windward and leeward surfaces of the through the gaps does not completely penetrate the dominant
panels, mostly in the gap region. It can be observed that the pres- vortex behind the panels. Therefore, by introduction of gaps in
ence of the gap leads to a reduction of pressure on the upper half the range of 0.0 to 0.2 m for this specific solar panel set, the flow
of the windward face of the lower panel, whereas it causes the over and around panels does not split into two parts behind the
pressure to drop along the entire leeward face of the lower panel. panels (i.e., gaps do not play the role of a splitter plate). Figures 7d,
For the upper panel, more pressure is exerted on the lower half of 7e, 7f illustrate that a symmetrical behaviour is observed in the x-z
the windward face, while the pressure on the leeward face is plane for the mean flow, which is in contrast with that observed in

Published by NRC Research Press


734 Can. J. Civ. Eng. Vol. 41, 2014

Fig. 5. Mean static pressure (Pa) distribution on (a) windward surfaces at ␪ = 0°, (b) leeward surfaces at ␪ = 0°, (c) windward surfaces at ␪ = 180°,
and (d) leeward surfaces at ␪ = 180°.
Can. J. Civ. Eng. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF CONNECTICUT on 10/11/14

Fig. 6. Pressure coefficient distribution over mid-section of the windward and leeward surfaces of panels with and without gap spacing.
For personal use only.

the x-y plane. Figures 7e, 7f also show that increasing the size of the observed in Cp curves (larger than no gap case) is due to the pres-
gap causes the small vortices to expand. This expansion leads to a ence of the additional small vortices formed in that region.
reduction in width and increase in length of the dominant vortex. Based on the analysis in the previous paragraph, the vortices
The flow structures shown in Fig. 7 can be used to explain the generated by the presence of the gap in the solar panel structure
flipping behaviour of the pressure coefficient curves for gap spac- significantly change the flow structure behind the panels. Al-
ing of 0.1 and 0.2 m in the leeward part of the lower panel, as though these vortices seem to be small, their repeated formation
observed in Fig. 6. As Fig. 7b shows (␦ = 0.1 m), there is a small and shedding can result in a fluctuating load on the panel. From a
vortex close to the leeward face of the lower panel that causes the structural point of view, in the long term, these fluctuations can
pressure to reduce over a wide area of the lower panel compared cause fatigue damage to the panels. It is apparent that some at-
to the case with no gap. For the portion of the leeward face where tention should be paid to the gap region while designing sets of
this vortex is not present, the Cp curve becomes similar to the solar panels with gaps, especially on how the gap affects the lower
behaviour of the no gap case. In Fig. 7c (␦ = 0.2 m), two small panels.
vortices can be observed in the wake of the lower panel. The small
vortex close to the gap generates a low-pressure region that can be 4.2. Ground clearance effect
observed in the Cp curve. The other reduction in pressure is due to Vortex shedding is the main reason for the occurrence of fluc-
the presence of another small vortex that forms at the bottom of tuating drag and lift forces on solar panels, which can cause
the leeward face. Generally, it should be noted that any reduction severe damage to the entire structure (Kopp et al. 2002).This

Published by NRC Research Press


Shademan et al. 735

Fig. 7. Sectional flow streamlines superimposed with the vorticity contours: (a, b, c) over mid-section of the windward and leeward surfaces of
panels for gap spacing 0.0 m, 0.1 m, 0.2 m, respectively; (d, e, f) in the middle of the panels in upper row for gap spacing 0.0 m, 0.1 m, 0.2 m,
respectively.
Can. J. Civ. Eng. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF CONNECTICUT on 10/11/14
For personal use only.

phenomenon is caused by the interaction of separating shear lay- for small ground clearance a complete suppression of shedding
ers, which results in a periodic flow regime downstream of the can occur. Suppression can be caused by the vorticity produced
panels. In proximity to the ground, the flow pattern is altered along the wall cancelling the vorticity in the lower shear layer of
since the lower shear layer is either weakened or completely sup- the panel (Durao et al. 1991; Wu and Martinuzzi 1997).
pressed (depending on the ground clearance of the panel) and the Three-dimensional RANS simulations have been carried out to
two separated flows tend not to interact in the downstream re- evaluate the effect of ground clearance on the wind loading ex-
gion. While vortex shedding exists for a larger ground clearance, erted on the entire structure of solar panels. Different typical

Published by NRC Research Press


736 Can. J. Civ. Eng. Vol. 41, 2014

Fig. 8. Pressure distribution over panels at different ground clearance with gap spacing of 0.2 m (␪ = 0°).
Can. J. Civ. Eng. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF CONNECTICUT on 10/11/14

installation heights were chosen (Hp = 0.5, 1.5, 2.5 m) to analyze 5.1. Wind loading and sheltering effect
the potential occurrence of vortex shedding downstream of the Figures 9a, 9b demonstrate the drag and lift force coefficients
panels. The solid area of the entire set is kept constant in all cases experienced by solar panels located in rows 1, 2, 3, and 4 with
and a gap spacing of 0.2 m was considered between each panel in different spacing ratios (S*/⌬). There is no significant difference in
the set. All assumptions made in the wind direction analysis (Sec- the wind loading of panels located in the front row (#1) at different
tion 3.2), including the boundary conditions and wall treatment, S*/⌬ ratios. Considering the second row of panels, the magnitude
For personal use only.

were considered here as well. It was shown earlier that the maxi- of the aerodynamic forces drops dramatically due to the shelter-
mum wind loading occurs at ␪ = 0° and 180°. Therefore, ␪ = 0° was ing effect caused by the panels in the first row. An increasing
selected for the ground clearance evaluation. trend is observed for the forces on the panels located in the third
Figure 8 compares the pressure distribution on the mid-section and fourth rows compared to the panels in the second row. These
of both the windward (upper curves) and leeward (lower curves) figures indicate that retaining a spacing ratio of S*/⌬ around one is
surfaces (either left or right column, see Fig. 1a) of the solar panels beneficial for reducing the loads on the panels in the array.
at different distances from the ground (Hp = 0.5, 1.5, 2.5 m). This To extract the reason for the reduction observed in the forces on
figure confirms the increase of the area under the pressure distri- the panels located in the second, third, and fourth rows, the pres-
bution curve as the distance from the ground is increased. It can sure distribution superimposed with the sectional streamlines of
be observed that increasing Hp does not have a significant influ- flow are plotted in Figs. 10a, 10b, 10c for spacings S*/⌬ = 1, 2, 3,
ence on the windward surface. However, it clearly reduces the respectively. As the pressure contours reveal, there is a maximum
pressure on the leeward surface, creating more aerodynamic pressure difference between the windward and leeward surfaces
force on the panels. of the panels in the first row relative to the other rows. At S*/⌬ = 1
The CFD results were compared with the data obtained from the (Fig. 10a), due to the short distance between the panels, the panels
ASCE 7-05 code for different ground clearance (not shown here). in the second row are located in the low pressure region in the
To use the ASCE code for force evaluation of solar panels, the wake of the first row. This produces an opposite direction for the
panels are considered as a mono-slope roof. The trend of the re- drag and lift forces on the second row at this specific S*/⌬ relative
sults between CFD and ASCE code was similar. It should be noted to the other rows of panels. This phenomenon can also be ob-
that the ASCE data represent the peak values of the wind loading served in Figs. 9a, 9b where the drag and lift coefficients of panels
and therefore an upward shift can be expected between the ASCE in row #2 have opposite signs compared to panels in the other
data and the CFD results, which represent the mean values. This rows. At the higher S*/⌬ ratios, the second row of panels is outside
analysis shows that increasing the ground clearance leads to an the low pressure region generated in the wake of the first row, and
increase in the aerodynamic force experienced on the solar panel. the pressure on the windward surface is increased. A similar phe-
nomenon is observed for the subsequent rows. This analysis indi-
5. Solar panels in arrayed configuration cates that solar panels in an arrayed configuration should be
Solar panels are widely used for residential and commercial carefully spaced to take advantage of the sheltering effect, since
purposes. In large power generation applications solar panels are the lift can be significantly reduced if the panels are properly
installed in an arrayed configuration (Fig. 1d). These arrays of pan- positioned.
els include a specific number of solar panels in different columns
and rows. The number of solar panels in each column and row is Conclusions
determined based on parameters such as the dimensions of the CFD simulations are used to predict the aerodynamic force dis-
installation field and the electricity generation requirements. tribution on a generic set of stand-alone solar panels consisting of
It is of interest to investigate the wind loading of solar panels in four individual panels in a 2×2 arrangement, and on four rows of
an arrayed configuration and also analyze the sheltering effect of solar panels in an arrayed configuration. Wind flow modeling
one row of panels on another. In addition, we aim to determine involving steady 3D RANS simulations was carried out using dif-
the spacing between solar panels which minimizes the wind load- ferent turbulence models and mesh topologies. Based on the val-
ing. Four sets of solar panels with different spacing between them idation studies, the final simulations were conducted using the
(S*/⌬ = 1, 2, 3) are considered. The 3D simulations are carried out at k−␻ SST turbulence model and a hybrid mesh scheme. Several
␪ = 0°, A = 135°, and Hp = 1.5 m. flow parameters, including the mean pressure coefficient, and the

Published by NRC Research Press


Shademan et al. 737

Fig. 9. Variation with spacing ratio of (a) drag and (b) lift. angles. Consequently, care should be taken to consider the lift and
moment when designing solar panel configurations. Considering
the complete structure, regions close to the bottom edge of each
individual solar panel experience the largest difference in pres-
sure between the front and back surfaces, making these regions
more vulnerable to wind loading.
Analysis of the gap spacing effect on wind loading exerted on
the solar panels shows that introducing the gap between the pan-
els significantly changes the flow structure in the wake region.
The size of the dominant vortex in the wake region reduces as the
Can. J. Civ. Eng. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF CONNECTICUT on 10/11/14

gap spacing is increased, causing a smaller pressure difference


between the windward and leeward surfaces. Small wake vortices
are generated close to the gap, which grow in size as the gap
spacing is increased. The presence of these vortices causes a local
drop in pressure thereby increasing the load on the panel in the
vicinity of the gap. Increase in the ground clearance causes a larger
pressure difference between the windward and leeward surfaces and
results in a larger mean wind loading on the structure.
Simulations of the flow over arrayed panels were performed for
spacing ratios ((distance between trailing edge of one row and
leading edge of the next)/(projected horizontal panel width)) of 1,
2, and 3. Results show that the smaller spacing ratio produces less
wind loading on the panels forming the array. Furthermore, at the
smallest ratio, the lift and drag forces on the panels located in the
second row have opposite direction relative to the larger spacings.
The second row of panels experiences the least amount of lift due
to the sheltering effect from the first row. Increasing the spacing
ratio increases the pressure difference between the windward and
leeward faces, consequently inducing larger mean wind loading
For personal use only.

on the panels in the array.


Due to the significant influence of unsteady wind loads on the
life span of solar panels, a suitable CFD simulation to evaluate the
Fig. 10. Pressure contours (Pa) superimposed with sectional vortex shedding and other flow characteristics should be carried
streamlines, (a) S*/⌬ = 1, (b) S*/⌬ = 2, and (c) S*/⌬ = 3. out to further improve the design of solar panels.

Acknowledgement
This research was supported by the Natural Sciences and Engi-
neering Research Council of Canada (NSERC) through the Vanier
Canada Graduate Scholarship program. Shared Hierarchical Aca-
demic Research Computing Network (SHARCNET) facility was
used for conducting the simulations.

References
ASCE 7-05. 2005. Minimum Design Loads for Buildings and Other Structures.
American Society of Civil Engineers, ASCE, New York, USA.
Bitsuamlak, G.T., Dagnew, A.K., and Erwin, J. 2010. Evaluation of wind loads on
solar panel modules using CFD. 5th International Symposium on Computa-
tional Wind Engineering (CWE2010), Chapel Hill, NC, USA.
Blocken, B., Carmeliet, J., and Stathopoulos, T. 2007. CFD evaluation of wind
speed conditions in passages between buildings: effect of wall-function
roughness modifications on the atmospheric boundary layer flow. Journal of
Wind Engineering and Industrial Aerodynamics, 95(9–11): 941–962. doi:10.
1016/j.jweia.2007.01.013.
Durao, D.F.G., Gouveia, P.S.T., and Pereira, J.C.F. 1991. Velocity characteristics of
the flow around a square cross-section cylinder placed near a channel wall.
Experiments in Fluids, 11: 298–304. doi:10.1007/BF00211788.
ESDU. 1982. Strong Winds in the Atmospheric Boundary Layer, Part 1: Mean-
Hourly Wind Speeds. Engineering Science Data Unit Number 82026.
ESDU. 1983. Strong Winds in the Atmospheric Boundary Layer, Part 2: Discrete
Gust Speeds. Engineering Science Data Unit Number 83045.
Fage, A., and Johansen, F.C. 1927. On the flow of air behind an inclined flat plate
of infinite span. Proceedings of the Royal Society of London, 116(773): 170–
197. doi:10.1098/rspa.1927.0130.
drag and lift forces at various wind directions, are compared with FLUENT 6.3. 2006. User’s guide. FLUENT Inc., Lebanon, N.H., USA.
Franke, J., Hellsten, A., Schlünzen, H., and Carissimo, B. 2007. COST. Best Prac-
the available experimental data. tice Guideline for the CFD Simulation of Flows in the Urban Environment.
For the stand-alone cases it is shown that the entire structure Action 732, Quality Assurance and Improvement of Micro-Scale Meteorolog-
experiences the maximum aerodynamic force at wind directions ical Models.
of 0° and 180°. At a wind direction of 0°, the bottom panels expe- Jubayer, C.M., and Hangan, H. 2012. Numerical simulation of wind loading on
photovoltaic panels. 43rd Structures Congress, Chicago, Il, USA.
rience the maximum force, while at 180°, the situation is reversed. Jubayer, C.M., Ayo, A., Siddiqui, K., Bitsuamlak, G., and Hangan, H. 2012. Numer-
Results confirm that the lift and moment are the dominant as- ical and experimental study of wind effects on photovoltaic (PV) panels.
pects of the aerodynamic load over a wide range of wind incidence AAWE12 Workshop, Hyannis, MA, USA.

Published by NRC Research Press


738 Can. J. Civ. Eng. Vol. 41, 2014

Kopp, G.A., Surry, D., and Chen, K. 2002. Wind loads on a solar array. Wind and Shih, T.H., Liou, W.W., Shabbir, A., Yang, Z., and Zhu, J. 1995. A new k - ␧ eddy-
Structures, 5: 393–406. doi:10.12989/was.2002.5.5.393. viscosity model for high Reynolds number turbulent flows - model devel-
Menter, F.R. 1994. Two-equation eddy-viscosity turbulence models for engineer- opment and validation. Computers and Fluids, 24: 227–238. doi:10.1016/
ing applications. AIAA Journal, 32(8): 1598–1605. doi:10.2514/3.12149. 0045-7930(94)00032-T.
Mohapatra, M. 2011. Wind Tunnel Investigation of Wind Load on a Ground Wolfshtein, M. 1969. The velocity and temperature distribution of one-
Mounted Photovoltaic Tracker. M.Sc. thesis, Colorado State University, CO, dimensional flow with turbulence augmentation and pressure gradient. In-
USA. ternational Journal of Heat and Mass Transfer, 12: 301–318. doi:10.1016/0017-
Patankar, S.V., and Spalding, D.B. 1972. A calculation procedure for heat, mass 9310(69)90012-X.
and momentum transfer in three-dimensional parabolic flows. International
Journal of Heat and Mass Transfer, 15: 1787–1806. doi:10.1016/0017-9310(72) Wu, K.C.Q., and Martinuzzi, R.J. 1997. Experimental study of the turbulent
90054-3. wake flow behind a square cylinder near a wall. Proceedings of ASME FED
Shademan, M., and Hangan, H. 2009. Wind loading on solar panels at different Summer Meeting, Paper FEDSM97–3151, Vancouver, B.C., Canada.
inclination angles.11th Conference of American Society of Wind Engineers, Wu, Z., Gong, B., Wang, Z., Li, Z., and Zang, C. 2010. An experimental and
Can. J. Civ. Eng. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF CONNECTICUT on 10/11/14

San Juan, Puerto Rico. numerical study of the gap effect on wind load on heliostat. Renewable
Shademan, M., and Hangan, H. 2010. Wind loading on solar panels at different Energy, 35(4): 797–806. doi:10.1016/j.renene.2009.09.009.
azimuthal and inclination angles. 5th International Symposium on Compu-
tational Wind Engineering (CWE2010), Chapel Hill, NC, USA.
For personal use only.

Published by NRC Research Press

You might also like