You are on page 1of 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/311858790

Characterization of dust accumulated on photovoltaic panels in Doha, Qatar

Article  in  Solar Energy · January 2017


DOI: 10.1016/j.solener.2016.11.053

CITATIONS READS

34 636

4 authors:

Wasim Javed Wubulikasimu Yiming


Texas A&M University at Qatar Texas A&M University at Qatar
30 PUBLICATIONS   199 CITATIONS    22 PUBLICATIONS   102 CITATIONS   

SEE PROFILE SEE PROFILE

Ben Figgis Bing Guo


Qatar Foundation Texas A&M University at Qatar
41 PUBLICATIONS   210 CITATIONS    64 PUBLICATIONS   1,376 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Photovoltaics Soiling - Field Measurement, Modeling and Mitigation View project

Energy Grand Challenge/QEERI View project

All content following this page was uploaded by Wasim Javed on 24 December 2016.

The user has requested enhancement of the downloaded file.


Solar Energy 142 (2017) 123–135

Contents lists available at ScienceDirect

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

Characterization of dust accumulated on photovoltaic panels in Doha,


Qatar
Wasim Javed a, Yiming Wubulikasimu a, Benjamin Figgis b,c, Bing Guo a,⇑
a
Texas A&M University at Qatar, PO Box 23874, Doha, Qatar
b
Qatar Environment & Energy Research Institute, HBKU, PO Box 5825, Doha, Qatar
c
ICube Laboratory, University of Strasbourg–CNRS, Strasbourg, France

a r t i c l e i n f o a b s t r a c t

Article history: In this study, samples of dust naturally accumulated for various exposure times on photovoltaic (PV) pan-
Received 7 June 2016 els were collected and characterized over a period of ten months in a solar test facility located in Doha,
Received in revised form 31 October 2016 Qatar. The dust accumulation rate (DAR) over the exposure time was determined gravimetrically. The
Accepted 29 November 2016
dust samples were characterized using particle size analysis, X-ray fluorescence (XRF), X-ray diffraction
(XRD), and scanning electron microscopy (SEM). The cleanness index change rate (CICR), a measure of
how fast the PV power output degrades due to soiling, was found to have strong negative correlation with
Keywords:
DAR, but the CICR/DAR ratio was found to differ between winter and summer. The DAR and the mean
Dust accumulation rate
Exposure time
particle size of the accumulated dust both decreased with increasing exposure time, reaching relatively
Particle size steady values for longer exposure times. Calcium was found to be the most abundant element in the
Chemical composition accumulated dust, followed by silicon, iron, magnesium and aluminum. Calcite, dolomite, and quartz
XRD were the dominant minerals in the accumulated dust, with gypsum being a minor component. Dust col-
XRF lected after dust-storm events had higher proportions of halite and quartz contents than non-dust-storm
days, depending on the direction of the wind. Also, dust particles accumulated on PV panels appeared to
agglomerate as the exposure time increased. The data provided in this paper will be useful for quantita-
tively determine the degree of soiling and its effect on PV performance in Qatar and regions with similar
environmental conditions. The data will also be useful for the selection of soiling mitigation technologies.
Ó 2016 Elsevier Ltd. All rights reserved.

1. Introduction efficiency (Darwish et al., 2015; Said and Walwil, 2014; Sayyah
et al., 2014; Tanesab et al., 2015). The effect of soiling is a function
Environmental concerns and growing energy demand have of dust loading on the PV module. Dust accumulation rate on mod-
increased interest in photovoltaic (PV) solar power worldwide in ule surfaces mostly depends on airborne particle concentration,
recent years, as a promising renewable energy source. Qatar, as distribution of aerodynamic particle size, and weather conditions,
well as other countries in the Middle East and North Africa (MENA) which are all site specific factors (Said and Walwil, 2014). The dust
region, has a tremendous potential for development and deploy- accumulation rate may also vary with the outdoor exposure time
ment of solar power generation, due to the high solar irradiation (Mastekbayeva and Kumar, 2000). For example, in the Greek capi-
levels and the availability of land. However, the extreme climatic tal, Athens, dust deposition masses of 0.1–1 g m2 were recorded
conditions and desert environment in the region pose significant for the outdoor exposure periods of 2–8 weeks (Kaldellis and
challenges to the successful deployment of solar power generation Kokala, 2010). A dust loading of 6.184 g m2 was reported for an
(Sarver et al., 2013). exposure period of ten months (February to December) in Dhahran,
Numerous studies have shown that dust accumulation on solar Saudi Arabia (Adinoyi and Said, 2013). In another work, a dust
surfaces can cause significant degradation of their solar conversion accumulation rate of 132 mg m2 day1 was found in Mesa, Ari-
zona (Boppana, 2015). Similarly, mass accumulation rates of 1–
Abbreviations: DAR, dust accumulation rate; CICR, cleanness index change rate; 50 mg m2 day1 in Colorado (Boyle et al., 2015) and 150–
PM10, particulate matter concentration with aerodynamic particle size 610 lm; 300 mg m2 day1 in the Minia region, Egypt (Hegazy, 2001), were
WS, wind speed; WD, wind direction; RH, relative humidity. observed depending on the tilt angle and location. More recently,
⇑ Corresponding author.
dust accumulation rates of 10–80 and 5–20 mg m2 day1 on 40°
E-mail address: bing.guo@qatar.tamu.edu (B. Guo).

http://dx.doi.org/10.1016/j.solener.2016.11.053
0038-092X/Ó 2016 Elsevier Ltd. All rights reserved.
124 W. Javed et al. / Solar Energy 142 (2017) 123–135

inclined glass plates have been reported at Commerce City and the (1) quantify the rate of dust accumulation on PV panels for various
Erie site, respectively, in Colorado, US (Boyle et al., 2016). exposure periods, and (2) determine the physical and chemical
The characteristics of deposited dust have a significant impact properties of accumulated dust, and their correlation with environ-
on soiling-related PV performance degradation. Various laboratory mental variables, time of exposure and soiling-induced PV perfor-
studies have shown that the impact of dust soiling on PV energy mance loss.
yield is dependent on the dust physicochemical properties
(Appels et al., 2013; Jiang et al., 2011; John et al., 2016; Kaldellis 2. Methodology
and Fragos, 2011; Kaldellis et al., 2011; Kaldellis and Kapsali,
2011; Khatib et al., 2013; Sulaiman et al., 2011). These laboratory 2.1. Sampling of accumulated dust
studies have identified 15 types of dust pollutants, of which six
pollutant types (i.e. ash, limestone, red soil, calcium carbonate, sil- Sampling of dust accumulated on PV panels was carried out at
ica, and sand) are believed to have a greater effect than others on the Solar Test Facility (STF) (at latitude 25°190 32.6100 N and longi-
PV performance degradation at a given surface loading (Darwish tude 51°250 59.8300 E) located at Qatar Foundation, Doha, Qatar.
et al., 2015). However, most of these studies have used artificial Samples of accumulated dust were collected from the surface of
dust, and there has been little research regarding the physico- twelve panels of a PV array (CdTe thin film frameless, Pmax
chemical characteristics of naturally deposited dust in real-world 90 W, tilted at 22° and facing due South). Each panel is 1.2 m in
operating conditions. Soiling studies with adequate natural dust width, 0.6 m in height and 6.8 mm in thickness. A photo of the test
chemical composition information would be useful for under- PV array is shown in Fig. 1. Of the twelve panels, four panels were
standing the relation between soiling-induced effect and chemical used to collect daily samples, i.e., accumulated dust was collected
composition. every 24 h from these four panels. The 24-h dust samples were col-
Understanding physicochemical characteristics of accumulated lected from 11 January 2015 to 20 February 2015 (winter period),
dust is also essential for the development of soiling mitigation and from 1 June 2015 to 15 July 2015 (summer period). The other
technologies. Chemical composition, along with particle size distri- eight panels were used to collect two-week, one-month, two-
bution, determines how dust particles interact with the PV surface, month and six-month accumulated dust samples; two panels for
and hence affects the applicability and effectiveness of PV soiling each sampling frequency.
mitigation technologies (Horenstein et al., 2013; Johnson et al., The matrix of dust sampling from the PV panels is shown in
2005; Kazmerski et al., 2016; Mazumder et al., 2015). Various the- Fig. 2. The environmental variables, shown in Fig. 2, have no strong
ories suggest that, by rendering the PV module surface hydrophilic trend that might be constructed to have a role in causing the
or hydrophobic, it is possible to mitigate soiling. However, it has dependence on exposure time. Dust accumulated on the surface
been shown that the affinity of particles to PV surfaces is not only of the PV panels was scraped off with a rubber spatula and care-
dependent on the PV surface properties, but also the particle prop- fully collected in a polystyrene Petri dish. The spatula used for
erties. These aspects highlight the importance of understanding scraping dust was made from a polyvinyl chloride acetate (PVCA)
the characteristics of deposited dust (Kim et al., 2016). There are card that measures 90 mm by 45 mm by 1 mm. A long side of
also active methods of PV soiling mitigation, through the mechan- the spatula was gently pressed against the PV panel surface and
ical or electrostatic removal of dust particles from the surface. moved down to scrape the dust down the lower edge of the panel.
Information on the particle properties will be useful for assessing The falling dust was collected in the Petri dish. Multiple scrapes
the adhesion force of the particles, or for determining their motion were used for each area of the PV panel until the area was clean;
in an electric field (Mazumder et al., 2015; Quesnel et al., 2015). i.e. there was no visible dust present. This procedure was carried
Soiling-related PV performance degradation is also dependent out for the entire PV panel so as the ensure all dust on the panel
on particle size distribution of dust deposited on the surface (El- was collected
Shobokshy and Hussein, 1993; Pulipaka et al., 2016; Said and
Walwil, 2014). Particle size plays a significant role in reflectance,
2.2. Analyses of the dust samples
scattering, and absorption of light incident on solar cells and in
turn, leads to PV performance degradation. The greater tendency
All collected dust samples were subjected to the analyses
of resuspension for larger particles, even at moderate wind speed,
described below. Daily collected dust samples during a month
promotes accumulation of smaller-size dust particles (Weber et al.,
were subjected to gravimetric analysis individually and then
2014). Finer particles may cause more significant performance
degradation of PV modules than larger particles for the same mass
of dust. Smaller particles, having the greater specific surface area,
may be distributed more uniformly as compared to the coarser
dust particles, thus reducing the voids between the particles
through which light can pass (Tanesab et al., 2015). Weber et al.
(2014) reported that particle sizes deposited on the PV surface
mostly lie within a range of 1–50 lm. However, overall there has
been limited research on particle size distribution of dust accumu-
lated on PV panels in desert environments.
Characterization of accumulated natural dust and its impact on
PV system efficiency are limited, given the fact that dust accumu-
lation is a complex phenomenon and varies with site-specific envi-
ronmental and weather conditions (Kazmerski et al., 2016; Mani
and Pillai, 2010). However, through systematically study dust
accumulation in various locations of interest to PV power genera-
tion, it is hoped that the PV soiling problem may be better under-
stood and addressed. This study was carried out in Doha, Qatar. It
was intended to characterize dust accumulation on PV surfaces Fig. 1. Photo of a test PV array used in this study. The 12 panels on the outside
through a systematic approach. Specifically, this study aimed to (perimeter) were used for dust sample collection.
W. Javed et al. / Solar Energy 142 (2017) 123–135 125

Fig. 2. PV Dust Accumulation Sampling Matrix and the environmental variables (Each strip represent an exposure period. Arrows indicate two dust-storm days of February 9
and June 23, 2015).

combined into one-month agglomerated samples for the other Change Rate (CICR), a measure of how fast the PV power output
analyses. However, daily samples collected on dust-storm days degrades due to soiling, was determined to quantify the PV soiling
(February 9 and June 23, 2015) were submitted for all analyses level. The value of the Cleanness Index ranges from 0 to 1, with 1
separately, without being combined with other daily samples. for a perfectly clean PV module. Details of the PV data collection
For gravimetric analysis, the Petri dish was weighed before and and Cleanness Index calculation have been described elsewhere
after collecting the dust sample to determine the dust sample (Guo et al., 2015). In addition, ambient dust concentration
mass, using a Discovery DV215CD semi-microbalance (Ohaus (PM10), wind speed (WS), wind direction (WD), and relative
Corp., Pine Brook, NJ, USA) with a sensitivity of 0.01 mg. The dust humidity (RH) were also measured at the site. PM10 concentration
samples collected in Petri dishes were acclimatized to the temper- was measured using a TSI 8533EP DustTrak DRX Aerosol Monitor
ature and humidity conditions in the laboratory (20–22 °C; 45 ± 5% (TSI Inc., Shoreview, MN, USA). WS was measured by a wind speed
RH) for at least 48 h before weighing. transmitter (Thies Clima, Gottingen, Germany), WD by a wind
Particle size distribution of the collected dust samples was mea- direction transmitter (Thies Clima, Gottingen, Germany), and air
sured using an LS13-321 MW laser diffraction particle size ana- temperature and RH by a hygro-thermo transmitter-compact
lyzer (Beckman Coulter Inc., Brea, CA) in the dry powder mode. (Thies Clima, Gottingen, Germany). The 24-h arithmetic mean val-
Mineralogy characterization was performed using an Ultima IV ues were calculated for PM10, WS, and RH. The 24-h mean of WD
X-ray powder diffractometer (Rigaku Corp., Tokyo, Japan) and inte- was computed by treating all angular measurements as points on
grated PDXL2 powder diffraction software. While a ZSX-Primus II the unit circle and computing the resultant vector of the unit vec-
X-ray fluorescence (XRF) spectrometer (Rigaku Corp., Tokyo, Japan) tors determined by data points (Guo et al., 2015).
was used for elemental analysis. A KratosAxisUltra, X-ray photo-
electron spectrometer, XPS (Kratos Analytical, Manchester, UK)
was utilized for surface compositional characteristics. Particle 3. Results and discussion
morphology analysis was carried out on a Quanta 400 scanning
electron microscope (FEI, Hillsboro, OR, USA) equipped with an 3.1. Dust accumulation rate
Apollo XP EDS system (EDAX Inc., Mahwah, NJ, USA). The XRD,
XRF, EDS, and XPS analysis were used to established and inter- 3.1.1. Seasonal variation of DAR
compare the compositional characteristics. The EDS and XPS give The daily average dust accumulation rate (DAR) along with the
the surface compositional information and were used to monitor daily average of four environmental variables, i.e., PM10, WS, RH,
the consistency in the results and complement the bulk informa- and WD, are shown in Fig. 3. It can be seen that daily DAR follows
tion from XRF. the same trend as the concentration of PM10, while on days when
WS was higher (except dust-storm events) DAR was reduced.
Mean values of DAR and environmental variables in the winter
2.3. PV performance and environmental variable measurements (11 Jan to 20 Feb 2015) and summer (01 Jun to 15 Jul 2015) periods
are given in Table 1. The DAR for the bi-monthly exposure was
Concurrent PV performance data were collected at the STF test- higher (142 mg m2 d1) in the winter period than in the summer
ing site. From the PV performance data, the Cleanness Index period (98 ± 15 mg m2 d1). The CICR for the bi-monthly exposure
126 W. Javed et al. / Solar Energy 142 (2017) 123–135

3.1.2. DAR dependence on exposure time


The DAR of PV modules also depends on the time of exposure to
outdoor environmental conditions. In this study, when examining
the relation between DAR and exposure time, we excluded periods
that contained rain events. Therefore, what is presented here is pri-
marily the effect of ‘‘dry” exposure time on DAR. However, in some
sampling periods, water condensation apparently occurred. We
observed signs of water streams flowing down the PV panel. Most
such observations were in the winter, when RH was high. The DAR
results are shown in Fig. 4. It was observed that recently-cleaned
panel surfaces retain more dust than dirty ones. DAR decreased
sharply during the first two weeks of exposure, and then became
relatively steady as exposure time increased.
DAR decreased with increasing time of exposure. For example,
the average DAR for the six-month exposure is 80 mg m2 d1
(14,547 mg m2 over 180 days), which is lower than the averages
for two-month (120 mg m2 d1), one-month (140 mg m2 d1),
two-week (155 mg m2 d1) and 24-h (260 mg m2 d1) exposure
periods. It can be seen that DAR decrease sharply during the first
two weeks of deposition and after that became relatively steady.
In other words, daily DAR was higher in the early stages of dust
accumulation and steadied as exposure time approached six-
months. The reason for this trend may be that as dust accumulates
Fig. 3. Daily (24-h) dust accumulation rate (DAR) and corresponding environmen- on the PV panels, the resuspension rate increases, and hence, the
tal variables during the winter (11-Jan to 20-Feb) and summer (1-June to 15-July) net dust accumulation rate decreases over time. Dust accumulation
sampling periods. as a function of exposure time for a short period (one month) has
been reported by Mastekbayeva and Kumar (2000). They also
observed a sharp decrease in DAR during the initial two weeks of
deposition, which dropped sharply from 440 mg m2 d1 during
(PV modules cleaned every two months) was also higher (1.1%
the first week to about 120 mg m2 d1 after two weeks, and after
d1) during the winter period as compared to the summer period
that, it became relatively steady up to one month of the study
(0.4% d1). Apparently, this was due to the lower WS and higher
period.
RH in the winter period, the favorable conditions for dust accumu-
The accumulated dust mass increased with exposure time as
lation on PV panels, even though PM10 concentration was lower in
shown in Fig. 5. On average, an accumulated dust mass of about
the winter period. As reported in our previous study (Guo et al.,
5 g m2 was recorded after a month-long accumulation, which
2015), the daily PV soiling rate (referred to as daily DCI) was pos-
then increased to 15 g m2 after six months of field exposure.
itively correlated with WS and negatively correlated with RH.
The cumulative dust mass over a particular period can be predicted
DAR results for various exposure periods from this study are
by fitting a power curve, as given in Fig. 5, where ‘‘D” is the expo-
compared with other reports, as shown in Table 2. The daily DAR
sure time measured in days. For example, one would be able to
in this study is in the same range as found in Bankok, Thailand
predict the total amount of accumulated dust after a year from
(Mastekbayeva and Kumar, 2000), Minia region, Egypt (Hegazy,
these equations (excluding the cleaning effect of rain). In this
2001), Rumah, Saudi Arabia (Jones et al., 2016) and Baghdad, Iraq
study, the amount of dust collected after a year would be
(Saidan et al., 2016). The DAR of our study is much higher than
32.87 g m2 (17.41 and 15.46 g m2 respectively for Winter and
Athens, Greece (Kaldellis and Kokala, 2010), Mexico City (Weber
Summer six-month periods), assuming that the weather conditions
et al., 2014), Mesa, Arizona (Boppana, 2015) and Commerce City
and all other factors are constant. It must be noted that the trends
and Erie, CO, US (Boyle et al., 2016). The various dust accumulation
may vary for different climatic conditions and geographical
results were recorded with widely varying exposure time, ambient
locations.
dust concentration, and weather conditions. Hence, there is wide
variability in the dust accumulation rates, showing its dependence
on the exposure period, time of the year, deposition surfaces, geo- 3.1.3. Correlation between DAR and other variables
graphical location and prevailing climate conditions (Said and The relationship between the DAR and airborne PM concentra-
Walwil, 2014). This variation highlights the need for measurement tion, as well as other environmental variables, is not commonly
of ambient dust concentration and weather conditions in PV soil- reported in the literature. Also, there is only scarce information
ing studies at a particular location, so that the soiling results may about the relationship between accumulated dust characteristics
be interpreted with the relevant input variables (environmental and PV performance reduction. Net dust accumulation on a PV
variables), and generalizable soiling relations may be quantita- panel is the result of simultaneous deposition of airborne dust
tively derived. and resuspension of deposited dust; the environmental variables

Table 1
Mean values of DAR and other variables.

Period Mean values (±STD)


Bi-monthly DAR (mg m2 d1) Bi-monthly CICR (% d1) CICR/DAR ratio (% per 100 mg m2) PM10 (lg m3) WS (m s1) RH (%) WD (°)
Winter 142 1.1 (±0.9) 0.7 140 (±50) 1.94 (±0.90) 62 (±5) 270 (±80)
Summer 98 (±15) 0.4 (±0.8) 0.4 153 (±60) 2.48 (±1.10) 36 (±10) 324 (±60)
W. Javed et al. / Solar Energy 142 (2017) 123–135 127

Table 2
Comparison of measured daily DAR of this study with other studies from various cities around the world.

Location Sampling period Deposition surface Exposure period Tilt angle, DAR Reference
orientation (mg m2 d1)
Doha, Qatar Jan-Oct 2015 PV modules Daily 22°, South 215–306 This study
(1.2  0.6 m2) Biweekly 155
Monthly 140
Bimonthly 120
Six-month 80
Bankok, Thailand Apr-May 1998 Plastic sheets 6 days- one month 15°, South 125–440 Mastekbayeva and Kumar
(30  30 cm2) (2000)
Minia region, 1999 Glass plates (7  7 cm2) Weekly 0°, Horizontal 230–330 Hegazy (2001)
Egypt 20°, South 150–200
Rumah, Saudi Jan-Dec 2015 Glass plates (5  5 cm2) Weekly for 12 months 0°, Horizontal 25–190 Jones et al. (2016)
Arabia 15°, South 35–165
Baghdad, Iraq Sept 2014 PV module Daily 30°, South-west 210 Saidan et al. (2016)
Weekly 57
Athens, Greece Aug-Sept 2009 PV modules 2–8 weeks 30°, South 15–20 Kaldellis and Kokala (2010)
(988  448 mm2)
Mexico City Dec 2012-May Glass plates (60 cm2) Weekly 0°, Horizontal 65 Weber et al. (2014)
2013 (24–102)
Colorado, US Aug 2011-Jun Glass plates 1–5 weeks 0°, Horizontal 20–50 Boyle et al. (2015)
2014 (10  10 cm2) 40°, South 12–35
Mesa, AZ, US 20–22 Oct 2014 PV module 2 days 33°, South-west 132 Boppana (2015)
Commerce City, Aug 2011-Jun Glass plates 1–5 weeks (mostly 0°, Horizontal 10–120 Boyle et al. (2016)
US 2014 (10  10 cm2) 2 weeks) 40°, South 10–80
Erie, CO, US Nov 2012-May 0°, Horizontal 5–60
2014 40°, South 5–20

Fig. 5. Cumulative dust loading increases with the exposure time in a power law
Fig. 4. The DAR decreases with exposure time in a ‘‘power law” manner. One two- manner.
month (February-March) dust sample was included in the winter, and two samples
(April-May and June -July) were included in the summer period. One six-month
sample (from March to Aug) was included in the summer period. The particle size analysis also shows that 35% of particles in the
daily-deposited dust are smaller than 10 lm.
The rate of resuspension is a function of RH and WS (Figgis
mostly influence have been reported to be WS and RH (Figgis et al., et al., 2016; Kim et al., 2016). Higher RH promotes high DAR by
2016). reducing the resuspension rate of dust particles. Overall, daily
The correlation between daily DAR and environmental variables DAR was greater on days with the higher RH levels. Higher RH pro-
for 24-h exposure time was also examined in this study. PM10 at motes adhesion of dust particles to PV surfaces and inhibits their
the site varied between 40 and 340 lg m3 with an average of resuspension by the wind (Kim et al., 2016). This aspect is consis-
146 ± 50 lg m3 during the sampling period. Relations between tent with the fact that higher RH increases dust accumulation onto
daily DAR and environmental variables are complex as the WS PV modules and hence promotes the adhesion of dust, as the water
and RH both have a dual role in overall dust accumulation on the content of the fallen particles forms a bonding force between the
PV surfaces. In general, the average daily DAR was higher at ele- particles and the PV surface. Such an effect of RH on PV soiling
vated PM10 concentration, higher RH, and lower WS. (Fig. 3). has been confirming in our previous study (Guo et al., 2015). The
PM10 has a less strong correlation (r = 0.41) with DAR that is due RH affects the threshold WS for resuspension of deposited dust
to the fact that particles larger than 10 lm in aerodynamic diame- particles. Ibrahim et al. (2004) demonstrated that the threshold
ter are deposited but are not counted in the PM10 measurement. WS to detach 50% of stainless steel microspheres (64–76 lm) from
128 W. Javed et al. / Solar Energy 142 (2017) 123–135

glass surfaces in a wind tunnel increased from 3.6 to 13.4 m s1 as difference in the CICR/DAR ratio. Further investigation is needed to
RH increased from 18 to 67%. This result suggests that the dust provide a satisfactory explanation for this difference.
resuspension effect of high-speed winds is enhanced when the Daily DAR has a stronger correlation with environmental vari-
deposited dust particles contain the little moisture, which makes ables than does CICR, according to the correlation coefficients in
them less sticky and more likely to be carried away by the wind. Table 3. DAR is a quantity that depends on PM10, WS, and RH,
When humidity is sufficiently high, condensation may occur on whereas the CICR depends directly on DAR (and hence indirectly
the accumulated dust (Figgis et al., 2016). There are two possible on PM10, WS, and RH) and apparently other variables that are yet
outcomes related to condensation. First, this may lead to cementa- to be determined. As reported in our previous study (Guo et al.,
tion, in which a crust forms between the particle and surface over 2015), CICR was also negatively correlated with ambient PM con-
successive cycles of condensation and drying (Sarver et al., 2013; centration, positively correlated with WS, and negatively corre-
Sayyah et al., 2014). Secondly, the condensation may lead to dew lated with RH. The data given in supplementary material also
formation and results in a partial cleaning effect (Caron and showed a strong correlation between monthly cumulative dust
Littmann, 2013). However, there is little information available loading and monthly CICR. Overall, daily PV performance was
about this cleaning effect. highly reduced on days with higher DAR levels. The data revealed
In contrast, high-speed winds cause higher resuspension of that the phenomenon of dust accumulation is extremely complex
already deposited dust on PV panels and hence decrease DAR. In and challenging to practically understand in the outdoor environ-
general, low-speed wind increases dust settlement, while high- ment, given all the factors that influence dust accumulation.
speed wind tends to remove dust from PV surfaces and results in
cleaning (Figgis et al., 2016; Mani and Pillai, 2010). However, on 3.2. Dust characteristics
some high wind-speed days, DAR was high. Those were dust-
storm events typically, when PM10 was very high. Nonetheless, 3.2.1. Particle size distribution
PV panel position (e.g., tilt angle, orientation) in relation to the The size distribution of accumulated dust particles depends on
wind also affects dust deposition. the exposure time of PV modules to the outdoor environmental
The correlation between DAR and WD is the weakest of correla- conditions. Mean particle size of the accumulated dust was found
tions examined in this study (Table 3). This might be explained by to decrease with increasing exposure time. The 90th percentile
the fact that the NW and NWW prevailing winds covered the entire (by volume) particle diameter of accumulated dust was 44, 36,
range of wind speeds, but winds from other directions typically 32 and 27 lm for 24-h, one-month, two-month and six-month
only occurred in the low WS range (Guo et al., 2015). Because long exposure times, respectively, as shown in Fig. 6. Deposited
DAR relatively strongly depended on WS, the peculiar WS, and dust for 24-h exposure has a higher proportion of larger particles
WD relation led to the weak correlation between DAR and WD (25–50 lm) and a median particle diameter of 17 lm, compared
(Guo et al., 2015). However, in locations where there is no such to dust accumulated for one-month exposure having median parti-
biased WS distribution, the dust accumulation rate can exhibit a cle size of 14 lm. For the six-month exposure, 90% (by volume) of
significant correlation with WD. A study carried out in Minia, Egypt accumulated dust consists of particles smaller than 27 lm and has
(Elminir et al., 2006) quantified dust deposition on the glass plates median, mean, and mode particle diameters of 9, 13, and 16 lm,
installed at different orientations and tilt angles, simultaneously. respectively.
The results indicate that at lower tilt angles, the dust deposition Month-by-month accumulated dust samples also show signifi-
was higher for glass samples facing the south and southwest, oppo- cant variation in particle size distribution (Fig. 7). It was found that
site to the prevailing wind direction (from the north and north- mean particle size of accumulated dust was smaller in April and
east). For north and northeast facing surfaces, the dust May. This might be due to the higher wind speeds and lowest rel-
accumulation was lower. ative humidity during these two months (data given in the Supple-
There is a relatively high correlation between daily DAR and mentary material). Mean particle size was larger in June and July,
CICR, as shown in Table 3. Of interest is the ratio of CICR to DAR, which had more dust storms. Typically, 90% particles (by volume)
which quantifies the PV performance degradation as a function of dust are smaller than 36 lm on average for all month-long dust
of dust loading on the PV surface. As can be seen in Table 1, the samples. The volume-weighted mean and median particle sizes of
CICR/DAR ratio of the winter period is higher than that of the sum- one-month accumulated samples vary within a broad range of 12–
mer period. The values of CICR/DAR ratio found in this study are 25 and 9–16 lm, respectively. It was also observed that about 35%
similar to the ratio of PV power output loss to dust loading as of daily-deposited dust and 40% of monthly-deposited dust com-
reported by other researchers in a laboratory study (El- prised particles smaller than 10 lm, which depicts the strong cor-
Shobokshy and Hussein, 1993). However, PV performance loss relation between DAR and airborne PM10 concentration (Table 3).
due to dust loading is affected by several factors, including particle Overall, the size analysis indicates that the longer the exposure
size, chemical composition and the angle of solar incidence (El- time (i.e., the older the dust accumulation) of the PV panel sur-
Shobokshy and Hussein, 1993; Zorrilla-Casanova et al., 2013). From faces, the higher proportion of smaller size particles as seen in
summer to winter, all these factors may vary and contribute to the the case of six-month long dust sample. Mean and median particle

Table 3
Pearson correlation coefficient matrix for daily DAR and CICR with the different environmental variables.

DAR CICR PM10 RH WS WD


DAR 1
CICR 0.60* 1
PM10 0.41* 0.15 1
RH 0.33* 0.26 0.17 1
WS 0.35* 0.22 0.19 0.39* 1
WD 0.09 0.08 0.07 0.37* 0.48* 1
*
Significant at p < 0.05 using 2-tailed t-test. The correlation analysis is based on the daily dust samples for 24-h exposure time and daily average values of other variables.
W. Javed et al. / Solar Energy 142 (2017) 123–135 129

mance degradation may be dependent on the particle size accumu-


lated. Finer particles may cause stronger degradation in PV
performance than larger particles, for the same mass loading
(Qasem et al., 2014). It is attributed to the fact that finer particles
have more specific surface area and distributed more uniformly as
compared to the coarser dust particles, thus reducing the voids
between the particles through which light can pass (El-
Shobokshy and Hussein, 1993; Tanesab et al., 2015). Smaller dust
particles scatter shorter wavelengths more than the bigger parti-
cles, result in more light attenuation, especially at shorter wave-
lengths, and this trend increases with accumulated dust mass
(Qasem et al., 2014). We anticipate that for particles of mixed size
distributions, smaller particles can fill the voids between large par-
ticles. Hence, the distribution becomes more packed, and further
reduces the available PV area for light capture.
In addition, the fundamental energy requirement for cleaning is
also dependent on particle size distribution of the deposited dust,
Fig. 6. Box plot showing particle size distribution of accumulated dust as a function
of exposure time. either mechanically or electrostatically. For example, Kawamoto
and Shibata (2015) reported a consistent drop in the dust removal
efficiency by an electrodynamic shield for smaller-size (<25 lm)
particles while larger (50–300 lm) particles are much easier to
remove. Other methods of cleaning are likely to encounter a simi-
lar decrease in cleaning efficiency as particle size distribution
decreases. For example, Appels et al. (2013) observed that bigger
dust particles (60 lm) seem to be more easily removed by rain
than smaller particles (2–10 lm). These impacts of particle sizes
highlight the importance of investigating the size characteristics
of deposited dust.

3.2.2. Dust chemical composition


The chemical composition of deposited dust has a significant
impact on particle adhesion with the glass surface and is conse-
quently essential for cleaning method designs, either mechanical,
electrodynamic, or the design of coating system (Horenstein
et al., 2013; Johnson et al., 2005; Kazmerski et al., 2016). Informa-
Fig. 7. Box plot showing particle size distribution of accumulated dust by month- tion on the chemical composition of dust will be useful for assess-
long exposure time.
ing optical properties of the particles, or for determining the
motion of the particles in an electric field (John et al., 2016;
size in the accumulated dust was found to decrease with increasing Mazumder et al., 2015). Knowledge of the chemical composition
exposure time, with the 90th percentile particle size falling from of dust may also help to identify its origin.
44 lm for 24-h exposure to 27 lm for six-month exposure. The XRD analysis of all 24-h and one-month accumulated dust sam-
observed smaller grain-size particles suggest that the dust is orig- ples show no significant variation in major mineral composition.
inated mainly from local active construction or vehicular sites with Typically, dust accumulated on modules for various exposure
minimal contribution of the desert sand, which has larger grains. times have a similar mineral composition, as illustrated in Fig. 8.
Particle size plays a significant role in both dust settling and It was determined that calcite (CaCO3), dolomite (CaMg[CO3]2)
resuspension of dust on a PV surface. The high tendency of resus- and quartz (SiO2) are the dominant minerals of all collected dust
pension of larger particles, even at moderate WS, results in a samples. Both calcite and dolomite compounds accounted for more
greater fraction of smaller-sized particles in the accumulated dust than 70% of the dust particle contents in all one-month accumu-
at longer exposure time. Boor et al. (2013) reported that multilayer lated samples (Fig. 9). The next most prevalent minerals are quartz
deposits show significantly higher resuspension than single layer and gypsum, with the remainder other primary silicate minerals
deposits. Furthermore, the bigger particles moving along the sur- (i.e., palygorskite, albite, cristobalite and kaolinite). Of these dom-
face after detachment strike stationary particles on the surface, inant minerals, six-month accumulated dust has a high proportion
inducing further detachment (Kim et al., 2016). Fine dust particles (15%) of gypsum (CaSO42H2O) compared to the 24-h, one-month
readily stick to the PV surfaces when there is no rain. This condi- and two-month accumulated samples (Fig. 8).
tion will be aggravated by high humidity to create a cementing High contents of calcite and dolomite in the dust mainly origi-
effect on the surfaces (Sayyah et al., 2014). Typically, we found that nate from local soils. Engelbrecht et al. (2009) reported that Qatar
95% (by volume) of accumulated dust particles lie in the 1–50 lm and UAE soils mostly contain significant amounts of calcite (33–
diameter range. The higher accumulation of smaller-size particles 48%) and dolomite minerals compared to other Gulf countries such
on tilted PV panels is due to the greater resuspension of larger par- as Iraq and Kuwait, where soil has substantial amounts of quartz
ticles from dry surfaces at higher WS. Other reports of outdoor and feldspar minerals. They also suggested that both Qatar and
experiments also show that particle sizes of deposited dust on the UAE receive wind-blown dust from the Arabian Peninsula, Iraq,
the PV surface mostly lie in the range of 1–50 lm (Weber et al., and Kuwait. Similarly, Kazmerski et al. (2016) analyzed the compo-
2014). sition of deposited dust gathered from module surfaces in desert
Particle size is also an important factor in the accumulated regions of Middle East, i.e., Libya, Dhahran, Riyadh, and Baghdad
dust’s ability to scatter and attenuate solar irradiation. Therefore, and found that the deposited dust is dominated by quartz, followed
for the same dust loading (mass) on the PV surface, PV perfor- by calcite and primary silicates (i.e., Feldspar and Muscovite). Aissa
130 W. Javed et al. / Solar Energy 142 (2017) 123–135

The dominant minerals, i.e., calcite and dolomite, are attributed to


construction activities in the local urban environment, as these are
associated with limestone, marble, Portland cement, and concrete.
Considering that Qatari soils are ‘‘Calcisols” and dominated by cal-
cite and dolomite minerals (Engelbrecht et al., 2009), wind erosion
from land and construction sites is likely to be the primary source
of dust deposited on PV module surfaces.
The dust particles elemental analysis by XRF for all collected
dust samples for various exposure times indicated that calcium
from calcite was the most abundant element, followed by silicon
from quartz and minor elements including iron, magnesium, alu-
minum, sodium and potassium (Fig. 10). All detected constituent
elements were matched to the corresponding phase identified by
XRD. The elements (Ca, Si, Fe, Mg and Al) are present in all dust
samples and mostly associated with natural soil crust, while
enriched in the coarse dust particles. The 24-h dust samples have
a relatively high proportion of Na and Cl elements originating
mainly from sea spray (Yilbas et al., 2015), compared to accumu-
lated dust samples associated with longer exposure times, suggest-
ing these elements are associated with large size particles. We
observed that 24-h accumulated dust has a high fraction of
larger-size particles. It is reported that mixing of dust particles
with sea salt systematically leads to the growth of particles and
dust particles become larger as more sea salt is mixed with them
(Zhang and Iwasaka, 2004). The sampling site is close to (approxi-
mately 10 km) to the Arabian Gulf so dust particles are likely
affected by sea salt. It was further evident that the 9-Feb dust
storm coming from the sea had very high fractions of these sea-
Fig. 8. Mineralogical composition by XRD analysis of accumulated dust as a associated elements.
function of exposure time. The six-month-long accumulated dust has relatively high con-
tent of sulfur (S), corresponding to gypsum, compared to dust sam-
et al. (2016) and Yilbas et al. (2015) have also reported the high ples of shorter exposure times (Fig. 10a). The S concentration can
content of calcite along with varying components of other silicate be associated with Ca in the dust as gypsum. Moreover, it has been
minerals in dust particles collected from various locations in the reported that the absorption of SO2 via heterogeneous reaction on
Middle East. So, it can be anticipated that dust mineralogy varies the mineral dust, particularly CaCO3, is an important sink for SO2
substantially for various locations and climate zones in the region. and contributes to sulfate particle formation (Zhao et al., 2013).

Fig. 9. Mineralogical composition by XRD analysis of monthly accumulated dust.


W. Javed et al. / Solar Energy 142 (2017) 123–135 131

Fig. 10. The comparison of the bulk elemental composition by XRF analysis of accumulated dust (a) by exposure time (b) by month.

So, it can be anticipated that deposited dust particles can age dur- exposed to high humidity, morning dew, or light rain and then
ing the long periods of exposure to SO2 in the atmosphere. In addi- upon drying formed a cementitious layer (Kazmerski et al., 2016;
tion to natural minerals, traces of contaminating elements (Zn, Cu, Sarver et al., 2013).
Ti, Ni, Cr) were also observed. The sampling site is located in an
urban area and surrounded by light vehicular traffic so may be 3.2.3. Dust particle morphology
affected by contamination sources such as the resuspension of road The SEM analysis indicated that dust particle morphology
dust particles, vehicle fleet exhaust, brake and tire wear. changes with increasing exposure time. The dust particles took
All collected dust samples were also analyzed by EDS and XPS to irregular shapes in shorter exposure time samples. In longer expo-
determine the surface chemistry of particles. EDS and XPS results sure time samples, particles tended to be agglomerated and cov-
(not shown) allowed us to compare and confirm the elemental ered by small features. Long-exposed dust samples had more
composition of the examined dust particles by XRF. The surface fine-size particles. The change in particle shape and size may result
analyses (EDS, XPS) did not show a substantial difference in ele- in a change in chemical composition of dust particles (Yilbas et al.,
mental composition of a dust sample when compared to the XRF 2015). Such particle morphology change with exposure time can
results. However, some differences were observed in S, K, Na, and be seen in Fig. 11, which shows the images for 24-h, one-month,
Cl concentrations, showing differences in bulk and surface concen- and six-month accumulated dust samples.
trations. The surface analysis revealed slightly higher concentra- The daily (24-h) dust-accumulated sample has bigger particles
tions of S, K, Na, and Cl, reflecting the association of these and individual particles have rougher and clearer edges. Particles
elements with particle surfaces. This suggests that these hygro- are uniformly distributed on the copper tape, covering the whole
scopic species, present on the dust particle surface, render the par- area without any identified particles cluster (Fig. 11a). The one-
ticle are more likely to adhere to the glass and bind together to month dust sample has more small size dust particles with a few
agglomerate or form cement like films. In such process, soluble bigger ones with the size of >20 lm, compared to 24-h deposited
materials on the surface of the particles are dissolved when samples during the same month of accumulation. In the case of
132 W. Javed et al. / Solar Energy 142 (2017) 123–135

Fig. 11. The SEM images of dust samples deposited for (a) 24-h, (b) one-month and (c) six-month exposure periods, at 200 (1st column), 800 (2nd column) and 1600 (3rd
column) magnification.

six-month accumulated dust sample, there is a predominance of moisture/humidity to be dissolved, which results in agglomera-
finer (<10 lm) particles which have a tendency to agglomerate. tion of grains into clusters (Kazmerski et al., 2016; Klugmann-
Most smaller particles are stuck together, forming numerous large Radziemska, 2015). The dissolution of these alkali and alkaline
clusters. The few bigger particles (with the size of 17–22 lm) are compounds results in the formation of a liquid solution (mud)
mostly covered with and surrounded by many smaller (<5 lm) between particles, and upon drying it adheres the particles by
particles, forming large agglomerates (Fig. 11c). bridging the gaps/cavities in-between the dust particles (Yilbas
The high proportion of smaller particles over an extended per- et al., 2015). This results in the cementation of dust particles with
iod of accumulation was due to the natural processes of resus- the PV module surfaces especially in hot-humid conditions like
pension by which larger size particles were removed by wind Doha. It is observed in the field that the six-month long exposed
or light rain. These results are in accordance with the earlier module has a thick cement-like dust layer that is firmly bound to
results of particle size distribution. It was observed that more the glass surface of the module. This cement film formed due to
particles appeared to be agglomerated and covered by small fea- the binding of the dust particles together as well as adhesion to
tures with longer accumulation time. It was reported that soluble the glass surface, which is tough to remove, and required exten-
material on the particles’ surfaces (like Na, Cl, K, S, C) reacts with sive scraping.
W. Javed et al. / Solar Energy 142 (2017) 123–135 133

Fig. 12. The comparison of particle size distribution of accumulated dust during the
normal days and dust-storm events.

3.3. Dust-storm episodes

Dust storms occur in this arid region of the Middle East. Dust
Fig. 13. The comparison of XRD patterns of the mineralogical composition of
storms severely cause a reduction of visibility, decline of solar radi- accumulated dust during the normal days and dust-storm events.
ation and, in consequence, reduce the output of solar devices. Eight
dust-storms (having daily mean PM10 concentration of
>200 lg m3) were recorded during the study period (January– During the dust storms, particle size of the dust accumulated on
October 2015). Of these, two are characterized and described in the PV panels was quite different compared to normal days
this section, one occurring in the winter period on 9-Feb and (Fig. 12). More precisely, large size particles (30–60 lm) were
another in the summer time on 23-Jun. The first dust-storm event dominant in the case of dust storms, unlike in normal conditions.
had daily mean dust (PM10) concentration of 207 lg m3 and the On the normal days, smaller particles are more prevalent and big-
second one 220 lg m3. Dust mass accumulated during these dust ger ones much rarer. Thus, particles deposited during dust storms
episode days was 762 and 266 mg m2, respectively. As expected, tended to be larger than those deposited on normal days.
the dust mass accumulated on PV surfaces during the dust-storm One of the most interesting findings of dust characterization is
days was significantly higher than on normal days. It was also that the chemical composition of dust particles deposited on PV
observed that the 9-Feb dust storm resulted in more dust accumu- panels during these dust-storm events was entirely different from
lation on PV panels than the other one of 23-Jun, as it was accom- normal days (Fig. 13 and 14). These XRF and XRD observations
panied by relatively the lower WS and higher RH - favorable were corroborated by the SEM-EDS composition (data not shown).
environmental conditions for the dust accumulation (data given Dust-storm particles have a high proportion of halite (NaCl) and
in the supplementary material). The corresponding cleanness quartz (SiO2) in addition to calcite and dolomite minerals, depend-
index change rate (CICR, a measure of PV performance due to soil- ing upon the wind direction of the dust-storm event and, in turn,
ing) was found as 2.41% and 0.64% per day during these dust the dust source. The dust storm on 9-Feb came from the Arabian
event days, respectively. Gulf (140°; the SE direction) and had about 25% halite (NaCl)

Fig. 14. The comparison of the bulk elemental composition by XRF analysis of accumulated dust during the normal days and dust-storm events.
134 W. Javed et al. / Solar Energy 142 (2017) 123–135

(originating from sea spray), and 13% quartz (SiO2). The dust storm Elminir, H.K., Ghitas, A.E., Hamid, R.H., El-Hussainy, F., Beheary, M.M., Abdel-
Moneim, K.M., 2006. Effect of dust on the transparent cover of solar collectors.
on 23-Jun came from the NW (315°) direction and had higher (21%)
Energy Convers. Manage. 47 (18–19), 3192–3203.
quartz mineral and 7% halite mineral, as shown in Fig. 13. So, it can Engelbrecht, J.P., McDonald, E.V., Gillies, J.A., Jayanty, R.K., Casuccio, G., Gertler, A.
be anticipated that dust source during these dust storms is not W., 2009. Characterizing mineral dusts and other aerosols from the Middle East
local. The source of the 23-Jun dust storm is probably the – Part 2: grab samples and re-suspensions. Inhal. Toxicol. 21 (4), 327–336.
Figgis, B., Ennaoui, A., Guo, B., Javed, W., Chen, E., 2016. Outdoor soiling microscope
Arabian desert that is mainly comprised of quartz and calcite for measuring particle deposition and resuspension. Sol. Energy 137, 158–164.
(Engelbrecht et al., 2009; Kazmerski et al., 2016). The 9-Feb dust Guo, B., Javed, W., Figgis, B., Mirza, T., 2015. Effect of dust and weather conditions on
storm most likely originated from the desert of UAE and showed photovoltaic performance in Doha, Qatar. In: First Workshop on Smart Grid and
Renewable Energy (SGRE). IEEE, pp. 1–6.
mixing of dust particles with sea spray that resulted in the increase Hegazy, A.A., 2001. Effect of dust accumulation on solar transmittance through glass
of sea-salt elements in the dust (Zhang and Iwasaka, 2004). covers of plate-type collectors. Renew. Energy 22 (4), 525–540.
Horenstein, M.N., Mazumder, M.K., Sumner, R.C., Stark, J., Abuhamed, T., Boxman, R.,
2013. Modeling of trajectories in an electrodynamic screen for obtaining
4. Conclusion maximum particle removal efficiency. IEEE Trans. Ind. Appl. 49 (2), 707–713.
Ibrahim, A.H., Dunn, P.F., Brach, R.M., 2004. Microparticle detachment from surfaces
exposed to turbulent air flow: effects of flow and particle deposition
The rate of dust accumulation, DAR, reaches approximately characteristics. J. Aerosol Sci. 35 (7), 805–821.
100 mg m2 d1 for a two-month exposure time, but its value Jiang, H., Lu, L., Sun, K., 2011. Experimental investigation of the impact of airborne
can be more than twice as high for short exposure times. The dust deposition on the performance of solar photovoltaic (PV) modules. Atmos.
Environ. 45 (25), 4299–4304.
90th percentile particle size (volume based) of the accumulated John, J.J., Warade, S., Tamizhmani, G., Kottantharayil, A., 2016. Study of soiling loss
dust is about 32 lm for two-month exposure, but this statistic gen- on photovoltaic modules with artificially deposited dust of different gravimetric
erally decreases with increasing exposure time. Calcium was found densities and compositions collected from different locations in India. IEEE J.
Photovolt. 6 (1), 236–243.
to be the most abundant element in the accumulated dust, fol- Johnson, C., Srirama, P., Sharma, R., Pruessner, K., Zhang, J., Mazumder, M., 2005.
lowed by silicon, iron, magnesium and aluminum. Calcite, dolo- Effect of particle size distribution on the performance of electrodynamic
mite, and quartz were the dominant minerals in all accumulated screens. In: Conference Record of the Industry Applications Conference, 2005.
Fourteenth IAS Annual Meeting. IEEE, pp. 341–345.
dust samples. Particle agglomeration occurs as the dust is allowed Jones, R.K., Baras, A., Al Saeeri, A., Al Qahtani, A., Al Amoudi, A.O., Al Shaya, Y.,
to accumulate on the PV panel. The dust characterization results Alodan, M., Al-Hsaien, S.A., 2016. Optimized cleaning cost and schedule based
will be helpful for understanding PV soiling in Qatar and regions on observed soiling conditions for photovoltaic plants in central Saudi Arabia.
IEEE J. Photovolt. 6 (3), 730–738.
that share a similarity in ambient dust and weather conditions.
Kaldellis, J.K., Fragos, P., 2011. Ash deposition impact on the energy performance of
The data contained in this paper will be useful for prediction of photovoltaic generators. J. Clean. Prod. 19 (4), 311–317.
PV performance loss as based on environmental conditions and Kaldellis, J.K., Fragos, P., Kapsali, M., 2011. Systematic experimental study of the
pollution deposition impact on the energy yield of photovoltaic installations.
development of optimum cleaning approach, which is of interest
Renew. Energy 36 (10), 2717–2724.
to project developers and operators. Kaldellis, J.K., Kapsali, M., 2011. Simulating the dust effect on the energy
performance of photovoltaic generators based on experimental
measurements. Energy 36 (8), 5154–5161.
Acknowledgments Kaldellis, J.K., Kokala, A., 2010. Quantifying the decrease of the photovoltaic panels’
energy yield due to phenomena of natural air pollution disposal. Energy 35 (12),
This work was financially supported by the Qatar National 4862–4869.
Kawamoto, H., Shibata, T., 2015. Electrostatic cleaning system for removal of sand
Research Fund (QNRF) through a NPRP grant (Project No. 7-987- from solar panels. J. Electrostat. 73, 65–70.
2-372) and a UREP grant (Project No. 15-083-2-030). Kazmerski, L.L., Diniz, A.S.A., Maia, C.B., Viana, M.M., Costa, S.C., Brito, P.P., Campos,
C.D., Neto, L.V.M., de Morais Hanriot, S., de Oliveira Cruz, L.R., 2016.
Fundamental studies of the adhesion of dust to PV module chemical and
Appendix A. Supplementary material physical relationships at the microscale. IEEE J. Photovolt. 6 (3), 719–729.
Khatib, T., Kazem, H., Sopian, K., Buttinger, F., Elmenreich, W., Albusaidi, A.S., 2013.
Effect of dust deposition on the performance of multi-crystalline photovoltaic
Supplementary data associated with this article can be found, in modules based on experimental measurements. Int. J. Renew. Energy Res. 3 (4),
the online version, at http://dx.doi.org/10.1016/j.solener.2016.11. 850–853.
053. Kim, Y., Wellum, G., Mello, K., Strawhecker, K.E., Thoms, R., Giaya, A., Wyslouzil, B.E.,
2016. Effects of relative humidity and particle and surface properties on particle
resuspension rates. Aerosol Sci. Technol. 50 (4), 339–352.
References Klugmann-Radziemska, E., 2015. Degradation of electrical performance of a
crystalline photovoltaic module due to dust deposition in northern Poland.
Renew. Energy 78, 418–426.
Adinoyi, M.J., Said, S.A.M., 2013. Effect of dust accumulation on the power outputs
Mani, M., Pillai, R., 2010. Impact of dust on solar photovoltaic (PV) performance:
of solar photovoltaic modules. Renew. Energy 60, 633–636.
research status, challenges and recommendations. Renew. Sustain. Energy Rev.
Aissa, B., Isaifan, R.J., Madhavan, V.E., Abdallah, A.A., 2016. Structural and physical
14 (9), 3124–3131.
properties of the dust particles in Qatar and their influence on the PV panel
Mastekbayeva, G., Kumar, S., 2000. Effect of dust on the transmittance of low
performance. Sci. Rep. 6, 31467.
density polyethylene glazing in a tropical climate. Sol. Energy 68 (2), 135–141.
Appels, R., Lefevre, B., Herteleer, B., Goverde, H., Beerten, A., Paesen, R., De Medts, K.,
Mazumder, M.K., Horenstein, M.N., Heiling, C., Stark, J.W., Sayyah, A., Yellowhair, J.,
Driesen, J., Poortmans, J., 2013. Effect of soiling on photovoltaic modules. Sol.
Raychowdhury, A., 2015. Environmental degradation of the optical surface of PV
Energy 96, 283–291.
modules and solar mirrors by soiling and high RH and mitigation methods for
Boor, B.E., Siegel, J.A., Novoselac, A., 2013. Monolayer and multilayer particle
minimizing energy yield losses. In: 42nd Photovoltaic Specialist Conference
deposits on hard surfaces: literature review and implications for particle
(PVSC). IEEE, pp. 1–6.
resuspension in the indoor environment. Aerosol Sci. Technol. 47 (8), 831–847.
Pulipaka, S., Mani, F., Kumar, R., 2016. Modeling of soiled PV module with neural
Boppana, S., 2015. Outdoor Soiling Loss Characterization and Statistical Risk
networks and regression using particle size composition. Sol. Energy 123, 116–
Analysis of Photovoltaic Power Plants. Arizona State University.
126.
Boyle, L., Flinchpaugh, H., Hannigan, M., 2016. Assessment of PM dry deposition on
Qasem, H., Betts, T.R., Müllejans, H., AlBusairi, H., Gottschalg, R., 2014. Dust-induced
solar energy harvesting systems: measurement–model comparison. Aerosol Sci.
shading on photovoltaic modules. Progr. Photovolt.: Res. Appl. 22 (2), 218–226.
Technol. 50 (4), 380–391.
Quesnel, D.J., Rimai, D.S., Schaefer, D.M., Beaudoin, S.P., Harrison, A., Hoss, D., Sweat,
Boyle, L., Flinchpaugh, H., Hannigan, M.P., 2015. Natural soiling of photovoltaic
M., Thomas, M., 2015. Aspects of particle adhesion and removal. In: Kohli, R.,
cover plates and the impact on transmission. Renew. Energy 77, 166–173.
Mittal, K.L. (Eds.), Developments in Surface Contamination and Cleaning,
Caron, J.R., Littmann, B., 2013. Direct monitoring of energy lost due to soiling on first
Fundamentals and Applied Aspects, vol. 1. Elsevier Inc., San Diego, CA, USA,
solar modules in California. IEEE J. Photovolt. 3 (1), 336–340.
pp. 119–145.
Darwish, Z.A., Kazem, H.A., Sopian, K., Al-Goul, M.A., Alawadhi, H., 2015. Effect of
Said, S.A.M., Walwil, H.M., 2014. Fundamental studies on dust fouling effects on PV
dust pollutant type on photovoltaic performance. Renew. Sustain. Energy Rev.
module performance. Sol. Energy 107, 328–337.
41, 735–744.
Saidan, M., Albaali, A.G., Alasis, E., Kaldellis, J.K., 2016. Experimental study on the
El-Shobokshy, M.S., Hussein, F.M., 1993. Effect of dust with different physical
effect of dust deposition on solar photovoltaic panels in desert environment.
properties on the performance of photovoltaic cells. Sol. Energy 51 (6), 505–
Renew. Energy 92, 499–505.
511.
W. Javed et al. / Solar Energy 142 (2017) 123–135 135

Sarver, T., Al-Qaraghuli, A., Kazmerski, L.L., 2013. A comprehensive review of the Yilbas, B.S., Ali, H., Khaled, M.M., Al-Aqeeli, N., Abu-Dheir, N., Varanasi, K.K., 2015.
impact of dust on the use of solar energy: history, investigations, results, Influence of dust and mud on the optical, chemical, and mechanical properties
literature, and mitigation approaches. Renew. Sustain. Energy Rev. 22, 698– of a PV protective glass. Sci. Rep. 5, 15833.
733. Zhang, D.Z., Iwasaka, Y., 2004. Size change of Asian dust particles caused by sea salt
Sayyah, A., Horenstein, M.N., Mazumder, M.K., 2014. Energy yield loss caused by interaction: measurements in southwestern Japan. Geophys. Res. Lett. 31 (15).
dust deposition on photovoltaic panels. Sol. Energy 107, 576–604. Zhao, Y., Chen, Z., Shen, X., Huang, D., 2013. Heterogeneous reactions of gaseous
Sulaiman, S.A., Hussain, H.H., Leh, N., Razali, M.S., 2011. Effects of dust on the hydrogen peroxide on pristine and acidic gas-processed calcium carbonate
performance of PV panels. World Acad. Sci. Eng. Technol. 58, 588–593. particles: effects of relative humidity and surface coverage of coating. Atmos.
Tanesab, J., Parlevliet, D., Whale, J., Urmee, T., Pryor, T., 2015. The contribution of Environ. 67, 63–72.
dust to performance degradation of PV modules in a temperate climate zone. Zorrilla-Casanova, J., Piliougine, M., Carretero, J., Bernaola-Galván, P., Carpena, P.,
Sol. Energy 120, 147–157. Mora-López, L., Sidrach-de-Cardona, M., 2013. Losses produced by soiling in the
Weber, B., Quinones, A., Almanza, R., Duran, M.D., 2014. Performance reduction of incoming radiation to photovoltaic modules. Progr. Photovolt.: Res. Appl. 21,
PV systems by dust deposition. Energy Procedia 57, 99–108. 790–796.

View publication stats

You might also like