Next Article in Journal
A Holistic Approach to Power Systems Using Innovative Machine Learning and System Dynamics
Next Article in Special Issue
A Review and Analysis of Green Energy and the Environmental Policies in South Asia
Previous Article in Journal
Global Gas and LNG Markets: Demand, Supply Dynamics, and Implications for the Future
Previous Article in Special Issue
A State-of-the-Art Review of Structural Testing of Tidal Turbine Blades
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Comprehensive Review of a Decade of Field PV Soiling Assessment in QEERI’s Outdoor Test Facility in Qatar: Learned Lessons and Recommendations

by
Brahim Aïssa
1,*,
Rima J. Isaifan
2,
Benjamin W. Figgis
1,
Amir A. Abdallah
1,
Dunia Bachour
1,
Daniel Perez-Astudillo
1,
Antonio Sanfilippo
1,
Juan Lopez-Garcia
1 and
Veronica Bermudez Benito
1
1
Qatar Environment and Energy Research Institute (QEERI), Hamad Bin Khalifa University (HBKU), Qatar Foundation, Doha P.O. Box 34110, Qatar
2
Division of Sustainable Development, College of Science and Engineering, Hamad Bin Khalifa University (HBKU), Qatar Foundation, Doha P.O. Box 34110, Qatar
*
Author to whom correspondence should be addressed.
Energies 2023, 16(13), 5224; https://doi.org/10.3390/en16135224
Submission received: 25 May 2023 / Revised: 30 June 2023 / Accepted: 3 July 2023 / Published: 7 July 2023
(This article belongs to the Collection Review Papers in Energy and Environment)

Abstract

:
Soiling of photovoltaic (PV) modules is a major issue due to its critical impact on PV performance and reliability, especially in the desert and arid regions such as the state of Qatar. Soiling frequently results in a severe reduction in PV power generation, which drastically affects the economical profitability of the PV plant, and therefore, must be mitigated. The most common way of mitigating PV soiling is surface cleaning. However, the latter could consequently increase the associated operation and maintenance (O&M) cost of the PV site. However, previous studies indicated that even if the best-optimized cleaning schemes are used, the actual global solar-power production can still be reduced by about 4%, which is associated with at least EUR 5 billion in annual revenue losses worldwide. This loss is expected to reach a conservative value of EUR 7 billion in 2023. Accordingly, investigating the interplayed physics phenomena related to the various soiling processes, the site-specific O&M costs, along with a techno-economical assessment of state-of-the-art soiling mitigation strategies (including innovative anti-soiling coating materials) is of paramount importance. The goal of this comprehensive report is to provide the solar community at large, and those focusing on the desert environment in particular, with real field measurements that provide key findings and challenges in addressing soiling research obtained from multiyear testing at the Outdoor Test Facility (OTF) field station, located in the desert environment of the city of Doha, in the state of Qatar.

1. Introduction

The term “soiling” may invoke different meanings. In the aerosol field, it is commonly attributed to the number of particles present on a given surface (e.g., surface density, which reflects the number or, in particular scenarios, the mass of dust particles per surface), while in the PV industry, it is mainly correlated with optical fixtures due to soiling accumulated on the front glass of PV module (e.g., light transmission, absorption, and reflection losses) as the absorbed light governs the associated PV module power directly. The focus here is to highlight the impact of the soiling on the PV energy yield generated in real-world conditions, such as outdoor test facilities (OTFs), such as the one located in Doha (Qatar). Subsequently, this impact might be discussed and compared with that observed in other geographical locations.
Recently, the topic of PV soiling has recorded a tremendous increase in terms of publications, as shown in Figure 1. The corresponding pie chart in the inset shows the associated fields placing “Engineering”, “Energy”, and “Material Sciences” as the most dominating topics. The investigation of the topic of PV soiling in the state of Qatar has been led by in-depth studies performed mainly in QEERI’s OTF and has resulted in multiple research works, especially during the period 2020–2022.
Within this report, the specific terminology is relevant. Hence, while the meaning of “soiling” is rather ambiguous, the terms “soil quantity” and/or “PV soiling loss” are used for their respective meaning. In fact, on a PV module, the decrease in light transmission (T%) due to dust accumulation does not perfectly match the associated PV power loss. More specifically, since solar cells have a selective spectral response, accumulated dust favors the attenuation of short wavelengths [1]. However, to avoid additional complexity, the attenuation of the transmitted light and the associated loss of the PV power are handled equivalently, unless stipulated otherwise, to avoid confusion. It was also reported that the correlation between PV power loss and dust accumulating mass had been occasionally found to be linearly proportional [2,3,4]. This is rather true during the first stages of the soiling process. Hence, once the surface becomes heavily soiled, additional dust particles cumulate onto the existing ones, therefore increasing the dust mass without necessarily affecting further the light transmission (i.e., flattening of T%) [5,6,7].
The geographical location also has a clear impact on the dust mass and, thus on the associated PV power loss. For instance, different types of dust (whether in terms of color, structure, size, etc.) have different impacts on light absorbance and/or transmittance. More specifically, for the same amount of dust, fine particles have a more dramatic effect on T% due to their higher cross-section (i.e., surface area to volume ratio). Similarly, the chemical composition and shape of the dust particles are critical factors that affect light absorption as well as light scattering properties [1]. Overall, the quantitative relationship between the cause and the effect (i.e., soiling mass vs. PV power loss) is linearly proportional for lightly soiled surfaces and nonlinear for heavily soiled ones (i.e., T% loss evolves more slowly than soil mass) [8,9]. Figure 2b shows a typical scenario of cleaned and non-cleaned PV modules after one month of outdoor exposure at the QEERI’s OTF in Doha, Qatar.
Succinctly, the soiling of PV surfaces relies on several parameters. First, it is impacted by site specificities, which include the location where the PV plant is installed, the associated concentration of airborne dust, the size of the particulate matter (PM), and the proportion of the soiling. Second, PV plant configuration and characteristics, which include PV module design, tilt angle, module’s orientation and height, the level of shading, and the cleaning frequency. Third, the meteorological variations and seasons such as ambient temperature, wind speed and direction, level of relative humidity, and the frequency of dust storms. Fourth, morphological, structural, and physical properties of the dust particles: including size distribution, dust potency, level of cementation/caking/capillary, aging, type of dust, color, density, etc. Fifth, surface properties of the front glass of the PV module, adhesion forces, and cementation.
This comprehensive review summarizes key field-measurement findings and challenges in addressing PV soiling obtained during the last decade at the QEERI’s OTF, located in the desert environment characterizing the state of Qatar. The manuscript is divided into ten distinct sections, each addressing different aspects of the subject matter. The introduction section presents the topic, provides background information describing the OTF, and highlights the significance of the review. Section 2 critically analyzes the impact of the soiling on solar radiation in Qatar, followed by the results related to the development of an in-house outdoor soiling microscope, in Section 3. In Section 4, the field measurements of PV soiling by commercially available sensors, including MarsTM and Dust IQ are detailed. Section 5 describes in detail the soiling properties in the desert environment, including physical characteristics, the fundamentals of soiling processes, the influence of environmental parameters, the interaction between dust particles and the surface of PV modules, the parameters influencing condensation, and the dew mitigation by heating. Section 6 emphasizes the impact of PV soiling on PV performance (e.g., the effect of PV module surface orientation; the dust potency of PV soiling loss, and the seasonal variability of PV soiling in Qatar), and Section 7 develops the mitigation of PV soiling (including topics related to the manual and automated cleaning, the development of anti-soiling coating and the soiling mitigation potential by1-axis PV trackers are developed). The effect of soiling on bifacial PV modules and the energy cost and power production are discussed in Section 8 and Section 9, respectively, and the renewable energy and soiling within the Gulf Cooperation Council (GCC) context is highlighted in Section 10. Lastly, the conclusion summarizes some key points discussed in the paper and offers a perspective on the topic, emphasizing its importance and potential impact.

Description of the OTF

Which solar-energy technologies work best in Qatar’s climate? In 2012, this question led to the establishment of the Solar Test Facility (STF)—updated now to Outdoor Test Facility (OTF), at Qatar Science & Technology Park (QSTP). The OTF is an outdoor site hosting about 90 PV systems from several manufacturers. Located at: Latitude 25.327044025° (19′37.36″ N); Longitude 51.432901051° (25′58.44″ E); Sea Level 9 m. It extends over a 35,000 m2 area and includes numerous types of supporting instruments for PV research, such as PV module test benches, trackers, and automated cleaning systems, among others. The OTF is continuously exposed to harsh environmental conditions, which makes it ideal to evaluate the efficiency of different PV systems under realistic conditions in Qatar. It is of paramount importance to evaluate the tested systems in terms of energy production and long-term reliability when exposed to elevated temperatures, high irradiation, intense humidity, and extreme soiling conditions, which all prevail throughout the year in the region (e.g., highest temperature: 42 °C, lowest temperature: 13 °C; Mean Temperature (Yearly): 27 °C; Mean RH (yearly): 56%; Precipitation: 0.2 mm; and Mean Wind speed (Yearly): 4 m/s). Koppen climate classification BWh (B = Arid; W = Desert; h = Hot arid).
A summary of typical equipment tested and installed at the OTF is the following:
-
38 crystalline silicon PV systems—mono-Si, multi-Si, bifacial
-
30 thin-film PV systems—CdTe, CIGS, μSi-aSi
-
4 hybrid PV systems—silicon hetero-junction, PERC, and TOPCon
-
7 concentrating PV systems
-
Inverters—central, string, micro-inverters
-
Trackers—1-axis, 2-axis
-
Battery storage system (500 kWh)
-
Linear Fresnel thermal collector
-
Anti-soiling coatings
Meteo station with a Global Horizontal Irradiance (GHI), Direct Normal Irradiation (DNI), Total Normal Irradiance (TNI), Diffuse Horizontal Irradiation DHI, Plane of Array (POA), UV, and albedo.
The full-scale testing commenced at the OTF in March 2013. In 2016, QEERI took over its operation. Since 2013, QEERI has started developing a unique regional and worldwide expertise on PV soiling in desert climates, both from experimental and modeling/forecasting points of view.

2. Impact of the Soiling on Solar Radiation in Qatar

The continuous assessment of solar resources is needed throughout the lifetime of solar-power projects, from the feasibility analysis and design to the setup and management of the system. In QEERI, solar radiation measurement, modeling, and forecasting have been developed as main capabilities to help Qatar in deploying solar-energy-based projects. For instance, QEERI runs a network of 13 high-quality solar radiation-monitoring stations across the state of Qatar, measuring the three components of solar radiation (direct, global or total, and diffuse), in collaboration with the Qatar Meteorological Department (QMD). Since Qatar is a desert area, research works on solar radiation in QEERI have included the study of atmospheric dust or the accumulation of dust on sensor surfaces.
In this section, some of the main findings are presented. To study the effect of the atmospheric constituents on solar radiation, mainly atmospheric dust, and aerosols, the use of a LiDAR ceilometer in the estimation of the extinction of the beam or DNI has been investigated [10]. DNI, which is the energy coming directly from the sun disk on a unit area normal to the sun rays, varies along the sun line direction due to changes in the atmospheric contents such as clouds, dust, etc. Hence, the relation between the ceilometer signal and DNI under cloud-free conditions was studied, i.e., in conditions where clouds are absent and only dust and aerosols are present in the atmosphere.
Moreover, the daily variation of the integrated backscatter and the hourly averages of DNI for all the selected clear days of one year at local noon were investigated. Higher values of the integrated backscatter were seen in the summer season and during periods of dust storms indicating higher aerosol loads. DNI did not show clear seasonal changes during the year where higher radiation is usually expected in the summer due to the seasonal change in solar position. It showed, however, a significant reduction in periods of dust storms as well as lower values during the summer. A clear relationship was found between the two measurements, relating to the hourly so-called DNI clearness index.
As demonstrated in many studies [11,12,13,14,15,16,17,18,19], atmospheric aerosols can cause the soiling of PV modules by dry deposition under arid conditions. To quantify the effect of aerosols on the attenuation of solar radiation from the high atmosphere to the ground level, the Aerosol Optical Depth (AOD) parameter was used. AOD is also known to be well correlated with the particulate matter (PM) mass concentration [11,12,13]. AOD can be derived from satellite images combined with chemical transport models [14]; however, ground measurements remain the best method to acquire data locally with low uncertainties. In QEERI, we investigated a method of deriving the AOD using spectral ground measurements based on a multi-filter rotating shadow band radiometer (MFRSR) [15,16]. The collected data were fitted with linear regression and only the data that were correlated with a good fit were kept for the analysis. AOD values were estimated for several days, as per the method described above. The values were compared with AOD derived from the Copernicus Atmosphere Monitoring Service (CAMS) [17,18]. A subsequent study, related to solar radiation and dust accumulation effect on pyranometers in Qatar, was also conducted [19].
Subsequently, the soiling rate of pyranometers operating in Doha, Qatar, where atmospheric dust is abundant, and dust storms are frequent, was investigated. The experimental setup consisted of measuring GHI using two pyranometers: one acting as a reference with daily cleaning, and one test sensor with varying frequencies of cleaning. Figure 3 shows a sample period of the soling ratio quantified as the ratio of the daily averages of GHI measured by the soiled sensor to that by the reference sensor vs. the number of days since the last cleaning. The red line is the linear fit function quantifying the correlation between the soiling ratio and time, giving the value of the loss per day between 0.3% and 0.5% depending on the season. It is worth noting that more losses are seen in the summer when dust events are more common, and rainfall is practically absent in Qatar.

3. Development of an In-House “Outdoor Soiling Microscope”

Among the various climate factors, wind speed, relative humidity, and PM concentration are the key factors influencing the soiling rate of photovoltaics [8]. However, to reach a higher correlation between these parameters and the soiling ratio, the frequency of measurement was increased by designing and fabricating an efficient device, operating in outdoor conditions, and is capable of quantifying the soiling rate in real time.
This device is called an “outdoor soiling microscope” and consists of an adapted digital microscope, connected to a computer. It can measure every few seconds the following parameters:
  • detect the onset of condensation,
  • detect the vanishing of the condensation,
  • measure the deposit dust particles that are above 10 µm2 in size, and
  • measure their subsequent removal.
A glass microscope slide was fixed to the shroud protecting the front of the microscope. This system was then inverted in such a way that settled dust is visible through the slide by the microscope (Figure 4a). In doing so, images of ~0.9 µm/pixel were captured. A LED and a sheet of translucent paper were used to optimize the contrast of the dust particles (i.e., versus the background) and to provide continuous and uniform day and night lighting (Figure 4b, right panel). A program script was then developed to obtain images at a controlled time-interval.
The resolution of such a developed system was beyond expectations. It approached 2–3 pixels across, indicating that the detection of particles of about 2–3 µm in diameter was feasible. Moreover, the “outdoor soiling microscope” could detect condensation droplets that formed on the surface (Figure 4c), along with their life cycle, i.e., commencement (onset), development, and disappearance.
To better assess the impact of the environmental conditions, including wind speed, relative humidity, and airborne PM10 concentration on dust accumulation, and to point out the favorable conditions for natural cleaning, the data set between day hours (6:00 a.m. to 6:59 p.m.) and night were separated. The reason for this is that during the night, there was no net resuspension that took place (Figure 4d). Doing so, clear correlations were observed as follows:
(a)
Increased relative humidity leads to a decrease in net resuspension.
(b)
Increased PM10 concentration increases dust accumulation.
(c)
Increased wind speed favors the natural cleaning by particle resuspension, and is triggered typically at 3 m/s.
(d)
Increased wind speed above 4 m/s greatly favors the particles’ resuspension.
(e)
During the nighttime, no clear correlations could be claimed.

4. Field Measurement of PV Soiling by MarsTM and Dust IQ Sensors

To provide high-quality soiling monitoring with high precision and reliability, a novel and cost-effective class of soiling sensors has recently emerged, showing a minimum need for water, cleaning, or maintenance when operated in the field. Some of these newly developed devices are already marketable, and their detection principle is based on the optical characteristics of the dust particles. Two brands, namely Atonometrics’ Mars soiling sensor and Kipp & Zonen’s Dust IQ, are the pioneering commercial devices for PV soiling monitoring.
In the same vein as the ongoing work on the outdoor soiling microscope, we have demonstrated field-test data measurements of the Mars™ soiling sensor installed in two different geographical locations in Qatar (Figure 5), characterized by hot, sunny, and dusty conditions.
Field data showed high accuracy and very good correlations between the traditional soiling measurements using a cleaned/uncleaned reference cell and the Mars™ output, therefore demonstrating it field-test validity. Similarly, the same study was conducted with a commercial Dust IQ to detect daily soiling, and a high correlation between these sensors and a PV cell was demonstrated, especially during the cold season. The mismatch of the soiling ratio (ΔSR) measured from the two methods, namely the Mars™/DustIQ and the conventional clean/soiled reference cell pair method, was found to be less than 0.11% [21] as shown in Figure 6.

5. Understanding of the Soiling Properties in the Desert Environment

5.1. Physical Characteristics

The physical and chemical properties of soiling impact dust particles’ interactions with the PV surface, and the level of involvement of environmental factors. Hence, the physical properties, including the structural, optical, morphological, magnetic, and chemical composition of dust particles collected from PV panel surfaces installed in OTF were systematically investigated.
As shown in the SEM in Figure 7a, various particle sizes and morphologies could be observed; however, the mean size might be considered at around 20 μm with a dominant spherical-like shape, overall [22]. Two shape factors were noticed, one approaching ~1 for the small particles which were mainly spherical, and the second ones of about ~3.5, similar to that observed in Saudi Arabia [23] and African Sahara (~3) [24,25,26,27,28,29].
The chemical elemental composition of these desert dust particles was probed using EDS (Figure 7b,c) which revealed rather a non-uniform distribution and various concentrations of these compositions, within the same set of particles.
As observed in OTF, potassium traces could be associated with sea salt as the state of Qatar is a peninsula located in the Arabian Gulf. Sulfur can be associated with anhydrite or gypsum component (CaSO4), etc. [25,26]. The XRD analysis of these dust particles was displayed in Figure 7d and the phase chemical formula was summarized in [22]. The quantitative analysis led to some dominant composites, including akermanite (Ca2Mg(SiO7)), wuestite (FeO), sillimanite (Al2(SiO4)O), olivine (Mg2(SiO4)), and calcite (CaCO3) and quartz (SiO2) given the desert nature of the soils.
Some traces of titanium were also detected and were attributed to a geographical factor which is the proximity of the OTF to the urban zone, where Ti may originate from the resuspension of dust particles induced by heavy traffic (i.e., tires) (Figure 8) [30,31,32,33,34,35].
Figure 9 shows a comprehensive literature review of particles accumulating on outdoor surfaces [29,30,31,32,33,34,35].

5.2. Fundamentals of Soiling Processes

In this study, a deep and detailed overview of the fundamentals of soiling processes has been provided by investigating the interplay between the macroscale environmental parameters and the microscale properties of dust particles [35,36]. Figure 10 shows a set of the main influencing factors on the PV soiling phenomenon. The color code is as follows:
(a)
Macroscopic environmental factors in blue.
(b)
Microscopic environmental factors in green.
(c)
Manageable impacting factors are in orange. The latter includes PV module design, glass-surface properties, etc.
Although the focus was devoted to PV module surfaces, the same reasoning applies to concentrated solar-power (CSP) mirrors as well. Moreover, dust particle adhesion forces and how they vary with cementation, caking and capillary aging, dust removal, and accumulation mechanisms, and their correlations with environmental parameters, were discussed. This study may serve as a basic guide and a starting step towards optimizing the various developed mitigation strategies, including module cleaning, anti-magnetic, and anti-soiling coating processes.
The main findings suggest the following:
(a)
The concentration of airborne dust was found to be a key factor in forecasting the soiling rates over different geographical locations over medium to long periods [33]. For a shorter period, such as day-to-day variation, additional environmental parameters play a more important role. Moreover, three parameters were found to constitute the best PV soiling predictors, namely (i) the airborne dust concentration, (ii) the duration of the dry periods, and (iii) the rainfall frequency.
(b)
The adhesion forces between the flat glass surface and the dust particles were dominated by capillary forces in the presence of moisture, which may also prevent their resuspension by the wind.
(c)
van der Waals forces dominated the adhesion in the case of dry conditions, though gravity and electrostatic forces could be considered negligible.
(d)
Under windy conditions, when drag forces were present, rolling was the dominant detachment mechanism for particles.

5.2.1. Influence of Environmental Parameters

Considering environmental conditions with low rainfall events and no condensation/cementation phenomena, the wind will be the prevailing natural cleaning process. For the dust particles to be effectively detached from the surface, the hydrodynamic forces (in addition to the wall shear stress) must beat the adhesion forces. When present, rain may efficiently clean surfaces. The correlation between rain events and surface cleaning was investigated in reference [33], and this study demonstrated that a threshold of minimum precipitated water was necessary for effective cleaning. This threshold was dependent on the velocity of the droplet, dust composition, surface wettability, tilt angle, and dust adhesion forces, and varied from a daily minimum of 0.3 mm [33] to 5 mm [36,37], 7–8 mm [38] and even up to 20 mm [39].
It was demonstrated that environmental factors greatly influence the various mechanisms of dust particle deposition and its adhesion/removal properties.
These environmental factors, including temperature (T), relative humidity (RH), wind speed (WS), and the concentration of airborne dust, vary along the day following a regular pattern [40] as follows:
  • Ambient temperature: lower during the night period, higher during the daytime.
  • PV module temperature: lower during the night period, higher during daytime (at even higher and lower levels than ambient temperature).
  • Relative humidity (RH): higher during the night period, lower during the day. Please note that the temperature of the air governs the concentration of the water-vapor saturation.
  • Wind speed (WS): lower during the night period (rarely above 3 m/s) and higher levels during the day going above 5 m/s.
  • PM10 concentration: there are no established patterns as observed for other parameters. However, statistically, we record higher mean levels during the night period than during the day. The maximum values of PM10 were observed early morning (6–8 am).

5.2.2. Interaction between Dust Particle and PV Module Surface

The accumulation of dust and the associated soiling rate is mainly governed by the behavior of the dust particle in terms of its interaction with the PV surface, and more particularly, by its rebound and resuspension versus its deposition (this will be more detailed in Section 5.3). These interactions are subjected to the surface properties, removal forces (e.g., by wind), and the particle/surface adhesion at large. Figure 11 depicts the main important soiling mechanisms, while Figure 12 summarizes the typical particle adhesion forces.
Typically, the nature of the adhesion forces is surface’ property dependent, and has several forms:
(a)
Rough and smooth glass surface: van der Waals forces.
(b)
Hydrophobic and hydrophilic surface: capillary forces.
(c)
Charged particles: Electrostatic forces.
(d)
Gravity, when assuming the weight of the particle sphere.
Properly addressing the type of interactions between the dust particle and the PV module surface is of paramount importance for two main reasons: (i) understanding the involved interactions and their correlation with the dust composition, and (ii) appropriate development of the most efficient cleaning method (dry/wet, robotic, etc.). In [41], the main fundamental adhesion forces, namely gravitational, van der Waal, capillary and electrostatic, were reported experimentally. These four forces dominate the early stages of the soiling mechanism. The findings demonstrated that under high relative humidity, the adhesion process is governed by capillary force, while during dry conditions, van der Waal force dominates (see Table 1). The main findings are detailed below:
Obtained results showed that the attraction mechanisms acting on the particle/surface under humid environmental conditions were 98% capillary, and 2% van der Waal, whereas the gravitational and electrostatic, were negligible. Please note that gravitational forces are often negligible for particles size below 500 µm [42,43,44,45,46,47,48,49].

5.3. Parameters Influencing Condensation

The investigation of the daily patterns of environmental parameters conducted in the QEERI/OTF site has demonstrated that relative humidity increased during nighttime, whereas wind speed and air temperature decreased (Figure 13) [49]. During the nighttime, and because of the radiative cooling phenomenon, PV modules’ temperature decreases (owing to their high IR emissivity) below the ambient one, which favors dew and water capillary formation, especially during the period just before dawn. In [50], experimental evidence has been provided demonstrating that the PV modules installed in OTF (Qatar) showed 5.1 K degrees below ambient temperature during the night.
The successive drying of the PV panel after the dew period increased dust particles’ adhesion, by triggering their cementation, caking, and capillary aging. In the latter scenario, mechanical cleaning was needed. According to this study, it was demonstrated that soiling occurs principally during the night period. Consequently, any cleaning (or self-cleaning) through wind may occur during midday time when the wind is at its maximum, and the surface is dry. However, this was true only for weakly attached particles that were not subjected to dew.
Experimentally, a naturally soiled glass coupon in the OTF field was still showing microscopic water droplets even when its temperature was superior to the dew point by about 14 °C. The water droplets were then growing when the gradient temperature passed to 8 °C and completely covered the glass surface when it was about ∼1–2 °C. However, while using a hydrophobic coupon, namely PTFE, microscopic water droplets were formed even when its surface was warmer than the dew point, but unlike the glass coupon, PTFE stopped the droplet growth and inhibited the flooding as the coupon was cooled. Figure 14 shows representative images taken by the developed outdoor soiling microscope (OSM). In ambient conditions, water droplets were formed on the surfaces of both glass and PTFE for 4 h, at nighttime.
Three processes govern the rate of dust accumulation onto a given surface (PV module or CSP mirrors), namely:
(a)
“Deposition” refers to the particle in the atmosphere impacting the surface.
(b)
“Rebound” is when this particle is rapidly rebounding from this surface without adhering.
(c)
“Resuspension” occurs once this particle is resuspended by wind [35,49].
Each of these processes may be signified as a dust flux rate, which translates the mass of the dust quantity per area and time units:
Accumulation = DepositionReboundResuspension
In [8,51], the accumulated soiling mass and the associated PV power loss were investigated along with the dust particle mechanics. Under desert environmental conditions in Qatar, dust particle sizes were between a few to tens of microns, while wind speed rarely exceeded a few meters/s. The main deposition mechanism was hence sedimentation, and consequently, the associated PV soiling was subjected to the tilt angle of the PV module. The sedimentation was more critical for larger particles. However, there was no universal model describing particle rebound, although resuspension was dependent on particle size (large particles were efficiently lifted off by the wind, unlike small ones) and was also favored by high wind speed as it was dependent on the square of flow (/wind) speed. Inclining the PV surface toward the airflow should promote the resuspension of particles, yet very few studies on the impact of the azimuth orientation (i.e., wind direction over the surface) on PV soiling were conducted, especially in real-world conditions. Thus, deeper research in this regard is necessary to study the mitigating of soiling by placing the surface of modules in different levels of airflow (e.g., by changing the height of the PV modules at elevated values). For more details, the reader was invited to consult references [8,49,51].

5.4. Dew Mitigation by Heating

Thermal treatment or heating the PV module to mitigate dew depositions can effectively reduce the cementation process [52,53]. Condensation usually occurs before sunrise when the ambient temperature is higher than the module temperature due to radiative cooling, and when the relative humidity is high. On days of dew prevalence, the soiling levels are recorded to be considerably higher than on dry days. Consequently, recent soiling reduction methods have been introduced by avoiding moisture through passive and active sheet heating. They may include thermal energy generated from thermal collectors (such as phase change materials) installed at the backside of PV models, active heating of PV modules using heat exchangers, or radiative cooling methods. Active heat with reasonably high energy suggested a soil reduction of up to 65%, but it seems that there are no data, simulations, or realistic assumptions on the financial viability of these warming strategies. These thermal systems can be retrofitted on PV modules and may serve dual advantages for devices; for daytime, cooling will result in increased energy yield whereas, for nigh time, heating will result in reduced condensation and hence cementation. The potential cleaning cost for implementation is less than 80 EUR/m2 for the PCM-based heating technique [53].

6. Impact of PV Soiling on PV Performance

6.1. Effect of PV Module Surface Orientation

One of the most critical parameters affecting the dust accumulation rate is the PV module orientation (Figure 15). This parameter is crucial too in the optimization of the overall PV performance. Commonly, only the sedimentation is taken into account, unless the conditions are rather windy with WS above ~3 m/s. Notably, the dust particle properties collected after a dust storm may differ from the dust particle properties collected on a normal day. The former will depend on the wind direction from where the dust storm is coming [54,55,56,57,58,59,60,61,62,63,64,65].
Sedimentation, which is affected by gravity, is dependent on the tilt angle, while inertial and resuspension, which are rather affected by wind, are dependent on both tilt angle and azimuth, the latter is the angle of the vertical axis of the PV module surface exposed to the horizontal airflow.
The accurate effect of the tilt angle on the soiling phenomenon is difficult to generalize because of the similarities between the deposition and resuspension trends (see Figure 10), and the competition between these two mechanisms is subject to the conditions specific to each situation.
Overall, results in desert environments show that deposition is more important than resuspension, and PV soiling is mainly controlled by sedimentation. In the experiments in wind-tunnel analysis, which tend to employ and investigate faster airflow speeds, it was found that soiling is more significant on moderately tilted surfaces rather than horizontal ones [65]. Ongoing research, initiated on glass coupons, is currently being conducted in OTF to identify the exact contribution of deposition and resuspension on PV modules as a function of the tilt angle.
In [61], we experimentally investigated the dust accumulation on glass coupons, positioned tilted perpendicularly to the wind. Two environments were studied, one based on a wind tunnel and the second representing the real-world conditions at the OTF site (Figure 16).
Parallel to this experimental investigation, computational fluid dynamics (CFD) modeling was performed to interpret the obtained experimental results. In OTF, it was found that:
(a)
The maximum dust deposition in the field test occurred at 45° facing the wind (this deposition was calculated as the total particles impacting the coupon surface).
(b)
The maximum accumulation occurred at 22° away from the wind (the accumulation was calculated as the total particles remaining on the coupon surface).
The difference between the two is due to the detached particles from the coupons when they were tilted toward the wind, which was dependent on two main parameters, namely the wind speed (wind flow velocity in this case), and the shear rate at the glass surface. The findings indicate that to decrease the soiling at night, a one-axis PV tracker (developed in Section 7.3) should be stowed at a maximum tilt angle, facing the wind.
Furthermore, to obtain more insights into soiling physics, the difference between dust deposition and its accumulation was studied. In this study, tilt angles were relative to the horizontal toward the wind direction. For instance, 45° indicates tilted facing the oncoming wind; 0° means horizontal; −45° implies tilted away from the oncoming wind (i.e., put downwind or backward, see Figure 16).
To simulate typical desert dust including that of Qatar, a dust-cloud producer dispersing “Belgian-Brabantian” dust, with 37 µm average size particle, was used in the wind-tunnel experiments [61,62,63,64,65]. However, as large dust particles tend to resuspend, mainly finer particles were found to accumulate on surfaces in this experiment [61,62,63,64,65]. At 5 m/s wind speed (which was the targeted speed in OTF), the average median diameter of the accumulated dust particles measured on the coupons was 20 µm, matching closely the 15 µm of the dust size collected from PV modules at OTF [8].
For the coupons that were facing the wind direction, the accumulation of the dust particles was found to decrease with respect to WS and with respect to the tilt angle too. Soiling reached its maximum of −22°, i.e., when the surface of the coupon was tilted slowly away from the wind. Moreover, tilting toward the wind was found to decrease the soiling as compared to the same tilt angle away from the wind. This indicates that for PV trackers that rotate commonly around ±45° or ±60°, a tangible reduction of dust accumulation could be achieved “for free,” especially during windy nights, by keeping the PV modules at a maximum tilt (ideally vertically) facing the wind.

6.2. Dust Potency of PV Soiling Loss

To assess the effect of dust deposition and accumulation on the PV panel surfaces, various techniques and approaches have been developed. Most of the studies available in the relevant literature discussed the performance degradation of the soiled PV modules during a given period of exposure [66,67,68]. However, only very few studies have been conducted on the quantification of the accumulated dust mass (ADM) effect on PV power loss [69,70]. The dust accumulation rate, at a particular location, was found to vary substantially with respect to the exposure time, season, and weather patterns [71,72]. Some works have been dedicated to depicting the experimental correlation between the ADM and the light transmission loss % [73,74,75], while others focused on the theoretical dependency between ADM and PV power loss [76].
In collaboration with Texas A&M University in Qatar (TAMUQ), a broader and deeper analysis approach to conclude an accurate quantitative relationship between dust-induced PV performance loss and the ADM and its physical and chemical properties was conducted [71]. Figure 17 summarizes the monthly average daily dust accumulation rate (DAR) patterns and weather and environmental variables that include airborne PM10, WS, and RH%, during the sampling period from January 2015 to October 2016, inclusively.
Depending on the dominant meteorological conditions, the monthly average DAR was found to vary from 75 to 250 mg/m2/day, with the lowest value observed during June and was attributed to the high WS, and low RH% during this month, although higher PM10 was observed during this month, which contributes to increasing the resuspension of the dust particles and therefore decreasing the net-accumulation of the dust on the PV surface [36,51]. This value of 75 mg/m2/day corresponded to a relative PV power loss of about 0.20%. Interestingly, the average daily DAR generally follows the same pattern as the airborne PM10 concentration, excluding the months with high WS where the DAR values are less. The findings indicate that higher DAR favors higher dust accumulation, and hence higher PV soiling was observed at elevated PM10 concentration, higher RH%, and lower WS parameters [71].

6.3. Seasonal Variability of PV Soiling in Qatar

The seasonal variation must be considered if one wants to forecast the soiling behavior, accurately estimate the PV power yield, and plan an optimized cleaning schedule. However, for a given geographical spot, a timescale of a single annual soiling rate might often be inadequate to precisely understand, and quantify the PV soiling, and the associated PV power loss given the temporal variability of the soiling ratio from one year to another [76,77,78]. Observing and identifying the seasonal soiling trends permits us to predict with better accuracy the associated total annual soiling losses. This is even true when optimizing the cleaning schedule. The quantification of the daily soiling and the accurate prediction of its rate before forecasting the seasonal soiling patterns and the corresponding average yearly energy loss is paramount in designing an economically profitable PV plant.
In an ongoing study with TAMUQ, we have in-depth studied the seasonal patterns of PV soiling in the desert conditions characterizing the state of Qatar, for six (6) years [73]. The field data related to PV power along with meteorological factors from 2014 to 2019 have been systematically and continuously recorded.
The cleanness index (CI) is defined as a measure of a PV module’s cleanness during a 24-h day. It is calculated as:
CI = PR T _ corr _ soiled PR T _ corr _ clean
where PR T _ corr _ clean is the daily temperature-corrected performance ratio of a clean PV array. PR T _ corr _ soiled is the quotidian temperature-corrected performance ratio of the PV array for which the cleanness is being assessed. The daily soiling rate is hence the daily variation of CI, i.e., “ΔCI” and is plotted in Figure 18.
For the day number n, the ΔCI which represents the difference of CI of the nth day and (n − 1)th one is:
Δ C I n = C I n C I n 1
Figure 19 shows the time series of soiling-induced PV performance loss in terms of CI for every month of the study period. CI, representing the PV soiling ratio, decreased substantially with respect to the exposure time. The soiled PV array was cleaned every second month (in addition to the natural cleaning by rain), which has restored the CI to unity.
  • The summer season (July–October) showed a decrease in CI by about 20% per month and this occurs consistently during only dry periods year after year.
  • Typically, for the bi-monthly cleaned panels, the associated PV power loss was about 15%/ month, with a clearly defined seasonal pattern.
  • During wet seasons, i.e., rainy periods, the CI reduction was limited to 10% due to cleaning triggered by the rain that prevents the accumulation of dust for more extended periods.
  • Wet season with higher rain rates such as in 2017 has shown a better CI (even close to unity) as compared, for example, to the same season in 2015.
  • Although the monthly PV soiling varied noticeably, the associated seasonal trends were more or less respected and the yearly change was less significant.
  • Cold, rainy, or warm seasons show different soiling as they are impacted by environmental factors.
  • A threshold of a minimum of 3 mm rainfall was found to be the required value to fully clean the PV modules.
  • Summer months are characterized by dust storms (DS), yet their impact when they occur during the winter season was found to be more impactful. DS days increase the annual average soiling rate by 23%.
This study highlighted the pattern of seasonal soiling in a desert environment characterizing the state of Qatar, which may help PV industrial operators in designing an efficient cleaning schedule and therefore results in a significant decrease in solar-power generation losses and the associated plant operational and maintenance costs, leading to a global revenue increase.

7. Mitigation of PV Soiling

7.1. Manual and Automated Cleaning

To date, there is no passive anti-soiling technology (based on surface coatings) that has demonstrated a complete elimination of the cleaning event. Moreover, no cleaning process is currently universally recommended, as it depends on the availability of on-site resources, the economics, and the required frequency of cleaning.
Generally, cleaning methods are classified into manual, semi-automatic, and fully automatic (Figure 20). An additional differentiation could be made between:
(a)
Dry-cleaning-based technologies: presently they are available only for PV and not CSP. These cleaning methods are applied in the desert and arid environments where water sources are rare,
(b)
Wet-cleaning-based technologies: these methods are usually favored owing to their efficiency and due to the low potential of surface damage [41,79,80,81,82].
Cleaning of the first mega-watt PV project in the region originally relied on manual cleaning; however, when robotic cleaning was introduced, the latter was faced with reliability issues due to the harsh soiling environment in the Arabian desert. Nevertheless, in the past three years, almost 100% of robotic cleaning was implemented in MW-scale PV projects. This has generated several questions regarding cleaning frequency, dust removal efficiency, and cleaning homogeneity over the whole PV surface area. Moreover, the concern of the possibility that PV surface coating shall be abraded by the robot movement was raised. All these concerns, whether technical, economic, or commercial, shall be further investigated to determine the best mode of cleaning where reliability and OPEX must be intact.
Nonetheless, for the present total solar power, the automated cleaning industry share is about 1.9 GW currently and represents only 0.13%. However, this is expected to expand to 6.1 GW by the end of the current year 2023 [40,83] thanks to the modern developments of dry and fully automated robots, which are increasingly implemented into the PV plant design (Figure 21).
A minimum of eight factors (Table 2) influence the decision on which optimal cleaning technology to adopt. This includes the nature of the soiling itself; the mass related to the dust deposition; the availability of water resources; the accessibility to the PV site; the configuration and architecture of the PV system, which includes tracking vs. fixed technologies, roof vs. ground-mounted configuration; the labor cost, the cost of the needed equipment and tools; and the conditions relative to the feed-in contract. Attempts are made to determine an appropriate cleaning frequency based on the identification of soiling levels and their forecasting. The potential cost for applying automated cleaning is in the range of 2.4–8.2 EUR/m2 with a possible soiling reduction of more than 95% [40,80,81,82,83,84,85]. The current ongoing efforts are focused now to identify an optimized cleaning schedule for Qatar using a machine learning forecasting model that is based on soiling-detection rate and local environmental conditions.

7.2. Development of Anti-Soiling Coating (ASC)

Anti-soiling coatings (ASC) is one of the leading soiling mitigation techniques thoroughly researched in the community. It is applied to the front glass of CSP mirrors and PV panels to minimize the soiling and decrease the need for frequent cleaning. Ideally, ASCs must be optically transparent, with antireflection and self-cleaning ability, non-toxic, stable, durable, cost-effective, and deployable at an industrial level. This rends ASC somehow the “Holy Grail” of the soiling community [86]. This passive cleaning strategy has seen minimal consumer adoption as it does not remove the necessity of cleaning but rather provides prolonged cleaning periods between successive cleaning events [87,88,89,90,91,92,93,94,95]. The soiling level in the field declined by as much as 80% using anti-soiling coatings. For extended times, nevertheless, typical anti-soiling efficiency is usually much weaker (for instance 20–50%) and may grow worse than bare front cover depending on the quality of coating, regional weather, and deterioration level. The major motivation, however, for anti-soiling costs comes from its potential cost which is less than 2 EUR/m2 [40]. Table 3 summarizes the state of the art of main PV soiling research and market.
However, limitations arising from physical phenomena must be considered:
(a)
Geographical and seasonal variation [96].
(b)
Particle adhesion physics [65,97,98,99,100].
(c)
Durability/Stability [101,102,103,104,105,106,107].
The long-term stability and reliability of passive coatings are difficult to foresee, and the PV-community has attempted to define standard methodologies to assess them a priori, e.g., VDI 3956-1 or IEC 62788-7-3.
Table 3. PV Soiling Research and Market—State of the Art.
Table 3. PV Soiling Research and Market—State of the Art.
A—International and Regional Efforts
InstitutionDetails of the Product
Regional
[108]KISR (Kuwait) Field Testing
[109]KAUST (KSA)Field Testing/Product Development—Cleaning Robot NOMADD
[110]DEWA (UAE)Field Testing
[8]QEERI (Qatar)Field Testing/Dust characterization/Fundamental Research/Anti-soiling coatings/Statistical Models
International
[111]FraunhoferPV Soiling and Degradation
[112]NRELPhotovoltaic Module Soiling Map, Forecasting Tools, Fundamental Research, Abrasion, Soiling, etc.
[113]AtaMoS-Tec—Chile project:
ISC Konstanz, Fraunhofer Chile, SERC and French CEA, INES
Photovoltaic Module Soiling study, Forecasting Tools, Fundamental Research
[114]SANDIA LABS/Arizona State University Soiling Loss Research, PV Reliability
[115]University of Colorado/Pontifícia Universidade Católica de Minas Gerais (Brasil)Soiling Science and Technology, Coatings and Films
[116]DLR Raumfahrtmanagement (the German Aerospace Center) Airborne soiling measurements and product development
[117]TÜV Rheinland (Germany)Anti-soiling coatings
[118]International PV Quality Assurance Task Force (PVQAT)Sensors and Monitoring, Cleaning Solutions and Anti-Reflective and Anti-Soiling Coatings, Standardization
[119]European Cooperation in Science and Technology (COST) “inDust” programInternational effort by WHO, WMO, ECMWF—Dust monitoring and forecasting models
B—Cleaning Solutions (Automatic, Robotic)
[109]NOMADD Desert, Utility Scale, Dry Brush
[120]Eccopia Desert, Utility Scale, Dry Brush
[121]WashpanelModerate Climate, Rooftops, Wet cleaning
[122]Greenbotics/ SunPowerGeneral Utility, Wet Cleaning
[123]First Solar/ DEWADesert, Utility Scale, Dry Brush
[124]Serbot GekkoModerate Climate, Rooftops, Wet cleaning
[125]SOLRIDERGeneral Utility, Wet Cleaning
[126]EnerwhereDesert, Rooftops,
[127]BladeRanger General Utility, Dry Cleaning
C—Soiling Sensors
[128]German Aerospace CenterQfly (Airborne soiling measurement of entire solar fields)
[129]Campbell ScientificSoiling Index Measurement Solution
[130]Kipp and Zonen DustIQ Soiling Monitoring System
[131]Nor-Cal ControlsMaxSun Soiling Station
[123]NRGSoiling Measurement Kit
[124]Ammonit Soiling Measurement Kit
[132]Atonometrics Mars Optical Soiling Sensor
[133]Kintech Engineering Soiling Measurement Kit
D—Smart PV monitoring systems (IoT/Data analytics)
[134]Alternative Energy SolutionsAES PIT (Uses machine learning/advanced data analysis platform)
[135]InnoEnergySolar Energy 3.0 (Smart PV monitoring esp. for detecting degradations)
[136]Solar IoT platformTrackSo (Data-driven predictive and condition monitoring)
E—Anti-Soiling Coatings
[137]CSD NanoMoreSun Multi-Function Coating (Electrodynamic Dust Shield, EDS)
[138]Anti-Soiling (AS) coatingDSM (Surface Modified Anti-Soiling Coating)
[139]Hydrophil AS coatingLotus Leaf Coatings (HydroPhillic Coatings)

7.3. Anti-Soiling Potential of 1-Axis PV Trackers

Expectedly, the surface tilt angle has a direct impact on soiling, where fewer dust particles may accumulate at steep angles, while more particles resuspend by tilting toward the wind direction, hence, the PV soiling can be efficiently reduced using a one-axis tracker stowed vertically at night, or by taking advantage from windy periods to help cleaning the PV modules by tilting the surface at angles that promote the particle resuspension. In [126], we experimented with the latter concept where a desert field study was conducted with 360° rotating 1-axis trackers and 10 cm × 10 cm glass coupons (Figure 22), with the trackers’ axis of rotation (East-West) approximately perpendicular to the dominant wind direction. It was found that a 41% soiling reduction was achieved on average by stowing vertically (90°) at night and a further 9% reduction by stowing upside-down (180°). However, attempts to promote dust removal by tilting coupons toward the wind during daytime windy periods achieved only around 5% soiling reduction (compared to the tracker remaining “on Sun”) and this approach would also sacrifice solar-energy harvest. In conclusion, modifying the 1-axis trackers to stow the PV modules at an angle ≥ 90° during the night period may efficiently reduce the PV soiling in the desert environment, while using these trackers to favor wind cleaning does not show to be an efficient process. Further validation of the findings is recommended using full-scale PV systems, in a variety of environmental conditions, especially where condensation is more prevalent and wind speeds higher.

8. Effect of Soiling on Bifacial PV Modules

It was reported that bifacial PV harvests sunlight from both sides of the module, which can significantly increase the energy yield per module by up to 30% compared to monofacial and therefore, the levelized cost of energy (LCOE). This bifacial ability is however subjected to many parameters including the height of the module, the distance between them, and the albedo. Studying the soiling effects in bifacial modules is critical since dust accumulation is one of the most impacting factors for PV power loss, and hence a relevant parameter to consider when designing the bifacial system.
Soiling on bifacial PV systems is a complex problem and requires an in-depth investigation of installation parameters: such as tilt angle, mounting height, and rear/front side dust deposition rates [127,128,129,130,131,132,133,134,135,136,137,138]. Vertical bifacial modules show great potential for negative soiling which can be employed as noise-barriers, building integrated systems, and for peak shaving of the power profile using east-west orientation [127,128,129,130,131,132,133,134,135,136,137,138]. In principle, all PV systems (bifacial or monofacial), experience a soiling effect; however, the effect on bifacial PV is less pronounced. This is primarily due to a smaller amount of dust exposure on the rear side of the module in tilted systems or due to the vertical installation practice of bifacial technology [138]. With the recent interest in bifacial PV installations, the impact of soiling on energy yield and LCOE has become the prime focus of the solar-energy community for accelerating its deployment [40,127,128,129,130,131,132,133,134,135,136,137,138].
To quantify the impact of vertical mounting of bifacial modules, Bhaduri and Kottantharayil [33] conducted energy loss experiments for three configurations: vertical bifacial (VB), latitude tilt bifacial (LB), and latitude tilt monofacial (LM) panels.
The experiments, conducted in Mumbai for 55.7% bifaciality factor, showed that vertical bifacial (VB) modules yield almost negligible soiling loss (0.027%/day) as shown in Figure 23a. On the contrary, latitude tilt bifacial (LB) resulted in a soiling loss of 0.39%/day loss and LM showed 0.40%/day as highlighted by the slope in Figure 23a. A similar work on the effect of tilt angle on a soiling deposition by Qasem et al. [1] also revealed that there was nine times less soiling on vertical modules than 30° tilted modules. For bifacial modules, the daily soiling rate was measured to be 0.236% whereas, for monofacial, it was 0.301%. Table 4 consolidates the results from bifacial/monofacial soling loss studies in the literature for comparison purposes.

9. Impact of Soiling on Energy Cost and Power Production

There is a clear competition between the cleaning cost and revenue losses originating from the soiling between the two cleaning events. To estimate the global impact and the associated cost of the soiling process with a descent accuracy, we have determined the optimum between these two competitive processes (i.e., soiling vs. cleaning) for the top 20 dominant PV and CSP markets. The associated dataset was compiled both from the relevant literature and exchanges with stakeholders. This study has included the soiling rates, cleaning costs, and simulated local energy yields, presented in Figure 24B, Figure 24C, and Figure 24D, respectively. Furthermore, the optimum number of the cleaning cycles was computed every year from 2017 to 2018 for each country, by considering the installed PV capacity as reported and the local feed-in-tariffs (Figure 24E), in addition to the medium growth scenario for 2023 and the average electricity price of about 0.03 EUR/kWh [40]. Moreover, the overall soiling costs were determined as the summation of the costs of the optimized yearly cleaning and the leftover income losses (Figure 24F). According to the data analysis, in 2018, the global solar-power production reduction to the soiling is estimated to be at least 4%, occasioning a conservative estimated monetary loss of at least EUR 5 billion [40]. Please note that additional costs related to optimized cleaning schedules (for instance in residences) may account for about 30% of global installations and cleaning rooftops was about eight times more expensive than cleaning ground-mounted PV [40]. Moreover, collateral effects such as increases in loan rates could have an additional financial impact. Thus, by 2023, the worldwide yearly power production losses due to soiling issues may rise substantially to 7%, generating more than EUR 7 billion [40]. Please note that in some regions of the world, additional specific parameters including air quality could reduce anthropogenic sources of soiling, albeit air-quality policies usually operate over long timescales [40].

10. Renewable Energy and Soiling within the Gulf Cooperation Council (GCC) Context

We conclude this comprehensive review by highlighting the effort deployed by GCC countries regarding renewable energy and the specific issue of PV soiling. More globally, the deployment of large PV plants in dusty environments such as the Middle East, North Africa, India, China, South America, and the US put a finger on the serious concerns of the associated O&M costs [139,140,141,142,143,144,145,146,147]. Although the MENA region has the biggest PV generation potential in the world thanks to the abundant sunlight, it is also the stronghold of harsh climate factors, especially the events of dust storms and high particulate matter concentrations. Soiling considerably decreases the light transmitted to the PV solar cells and hence lowers the output PV power.
W. Al-Kouz et al. [148] proposed computational models based on optimized architectures of artificial neural network (ANN) and extreme learning machine (ELM), using field data, therefore investigating the impact of dust accumulation and ambient temperature on the PV performance (specifically the power conversion efficiency) in Jordan. The developed ELM model was found to forecast the PCE with an accuracy translated by an R2 of 91.4%, thus proposing an optimized cleaning frequency every two weeks. Lopez-Lorente and coworkers [149] focused their study on Cyprus, where an outdoor soiling test-bench was assessed for a continuous period of two years. A seasonally dependent soiling rate ranging from 0.039 to 0.535%/day was pointed out. Furthermore, using ML approaches and physical models, six different models of soiling were studied, in terms of their performance (i.e., accuracy) and their limitations. Physical models demonstrated better performance than ML ones, and the study provided useful information on the performance and limitations of the different soiling models applied in dry and arid climates. To predict the soiling losses occurring on PV systems, Micheli et al. [150] studied 102 different environmental and meteorological parameters where the performances have been compared over 20 soiling stations installed over the USA. Findings first demonstrated that the yearly average of the daily mean PM value was the best parameter for soiling predictors. Second, the precipitation pattern was also pointed out as a very relevant predictor parameter among the other meteorological factors. A preliminary investigation of two-variable regressions resulted in an R2 of 90%. In recent work, Micheli and coworkers [150] have also employed satellite-derived or ground-mounted PM data and demonstrated that this can lead to mapping the soiling, in different geographical locations, with high accuracy. In 2023, S. Bhaduri et al. [151] investigated the effect of raindrops, acidic water, UV radiation, and abrasion on the durability of anti-dust coatings. The rainy season was found to degrade these ADCs by 21 times compared to the non-rainy one. Interestingly, all coated samples showed lower coating life by 10 to 48 times when exposed to the impact of raindrops than those exposed to water immersion (water kept at a pH of 7), therefore indicating the impact of raindrops hitting, causing more damage than a simple water contact. These studies also highlighted the alternative chemistries of hydrophobic and hydrophilic coatings that may be effective in exhibiting long-term reliability.
Notwithstanding this inhospitable climate, various projects of utility-scale PV plants have been successfully commissioned. Table 5 shows the estimated renewable energy capacity in GCC by 2030. Figure 25b summarizes the current targets, where UAE is leading, followed by KSA. Figure 25d summarizes the renewable power planned additions by country. Looking forward, the state of Qatar has set a goal of attaining 20% of its energy from solar power by 2030 (one of the largest PV plants in the world), namely Al-Kharsaah, with 800 MWp already commissioned, and two more projects starting in 2023 will account for 875 MWp and will be delivered by end of 2024. Led by the UAE, the GCC region has about 7 GW in renewable power generation capacity. Solar PV remains the dominant technology with more than 75% share, followed by 10% of CSP and a 9% share for wind projects.
Towards such an effort, GCC is rapidly increasing the achievement of utility-scale PV plants, such as Noor Abu Dhabi (1.2 GW), Al Dhafra (2 GW), and Al-Kharsaah (800 MW) plants. We are now welcoming the new terawatt era of photovoltaic (PV) solar energy. The world’s cumulative installed solar PV capacity grew by 22% to 940.0 GW by the end of 2021 claiming a 56% share of all renewable energies.
By May 2022, the installed capacity surpassed the milestone of 1 TW, led mainly by the deployment of large utility-scale solar plants. Although solar irradiance in the MENA region is among the highest in the world, soiling of solar collectors has been recognized as the main issue and the biggest detriment for solar-energy systems, decreasing their efficiency and increasing the cost of O&M.
Table 6 summarizes the main findings of the different sections developed during this review paper.

11. Conclusions

The utilization of renewable energy in an intensive carbon-producing world has become a necessity to abate climate change and greenhouse gas emission threats. Among the different available options, solar energy is considered the most promising solution in sunny areas.
This comprehensive review summarizes the experience with key field-measurement findings and challenges in addressing soiling research obtained from the last decade of testing at the Outdoor Test Facility. PV soiling has been demonstrated to be a complex phenomenon, with a high degree of freedom and various interplayed factors, including environmental parameters, but also physicochemical, structural, and morphological properties of the dust particles.
Understanding, mitigating, and forecasting PV soiling is still in its infancy, and extensive research efforts are needed to solve this issue. At a short time scale, robotic cleaning, optimized cleaning schedules and forecasting of the soiling phenomenon based on atmospheric factors inputs are the most attractive options. PV cleaning machines are now available in the market. At a longer time scale, passive anti-soiling technologies, including anti-dust coatings, could decrease the frequency of the cleaning events and thus the associated O&M cost. This is supported by the promising results that were already obtained with textured films, including porous silica, metal oxides, and fluorides; however, the reliability of the passive anti-dust coatings should be assessed further towards their large-scale and commercial deployment.
Through the developed expertise gained from exposing PV panels to the harsh environment characterizing the desert of the state of Qatar, QEERI is offering a unique opportunity to test, assess and create innovative mitigation technologies that will help GCC arid regions and the world. The anti-soiling coating may serve as an efficient solution that complements the active cleaning process. Indeed, innovative solutions, cleaning concepts, and novel coatings are continually developed, addressing new functionalities, including self-healing, condensation run-off, and retrofit applications. Finally, QEERI has recently launched an ambitious project for the Dust Atlas, which revolves around the understanding, forecasting, and mitigation of environmental conditions, to address this issue from all possible angles. Fifteen (15) meteorological and dust-sensing stations have been installed over the state of Qatar to monitor the real-time PV soiling both geographically and seasonally.

Funding

This research received no external funding.

Acknowledgments

The authors gratefully thank Bing Guo and his Team from the Mechanical Engineering Department, at Texas A&M University in Qatar, for their relevant input and fruitful exchanges on this topic. B.A. is grateful to A. Baloch for his support in drawing illustrative figures.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ΔSRSoiling ratioMENAMiddle East and North Africa
ADMAccumulated dust mass IRENAInternational Renewable Energy Agency
DSDust storms QEERIQatar Environment and Energy Research Institute
CICleanness index QMDQatar Meteorological Department
ARAntireflectionOTFoutdoor test facility
OPEXOperational expenditureGCCGulf Cooperation Council
PMParticulate MatterQSTPQatar Science & Technology Park
ASCAnti-soiling coatingTAMUQTexas A&M University in Qatar
DARDust accumulation rate SEMScanning Electron Microscopy
AODAerosol Optical DepthOSMOutdoor soiling microscope
O&MOperation & maintenanceXRDX-ray Diffraction
CFDComputational Fluid DynamicsSTFSolar Test Facility
CSPConcentrated Solar PowerkWKilowatt
DNIDirect Normal IrradiancePVPhotovoltaic
GHI Global Horizontal Irradiance HSATHorizontal single-axis trackers
DHI Diffuse Horizontal Irradiation PERCPassivated Emitter and Rear Cell
TNITotal Normal IrradianceTOPConTunnel oxide passivated contact
POAPlane of ArrayRHRelative humidity
LCOEThe levelized cost of energyWSWind speed
LEDLight emitting diodeWDWind direction
MFRSRMulti-filter rotating shadow band radiometerTTemperature
CAMSCopernicus Atmosphere Monitoring Service T%Transmittance
PTFEPolytetrafluoroethyleneIRInfrared
PCMPhase change MaterialVBVertical bifacial
LMLatitude tilt monofacialLBLatitude tilt bifacial

References

  1. Qasem, H.; Betts, T.; Müllejans, H.; AlBusairi, H.; Gottschalg, R. Dust Induced Shading on Photovoltaic Modules. Prog. Photovolt Res. Appl. 2014, 22, 218–226. [Google Scholar] [CrossRef] [Green Version]
  2. Sarver, T.; Al-Qaraghuli, A.; Kazmerski, L. Comprehensive Review of the Impact of Dust on the Use of Solar Energy: History, Investigations, Results, Literature, and Mitigation Approaches. Renew. Sustain. Energy Rev. 2013, 22, 698–733. [Google Scholar] [CrossRef]
  3. Sayyah, A.; Horenstein, M.N.; Mazumder, M.K. Energy yield loss caused by dust deposition on photovoltaic panels. Sol. Energy 2014, 107, 576–604. [Google Scholar] [CrossRef]
  4. Brown, K.; Narum, T.; Jing, N. Soiling Test Methods and their Use in Predicting Performance of Photovoltaic Modules in Soiling Environments. In Proceedings of the Conference Record of the IEEE Photovoltaic Specialists Conference, Austin, TX, USA, 3–8 June 2012. [Google Scholar] [CrossRef]
  5. Ahmad, Y.; Ghoneim, A. A new correlation between photovoltaic panel’s efficiency and amount of sand dust accumulated on their surface. Int. J. Sustain. Energy 2005, 24, 187–197. [Google Scholar]
  6. Molki, A. Dust affects solar-cell efficiency. Phys. Educ. 2010, 45, 456. [Google Scholar] [CrossRef]
  7. Elshobokshy, M.; Hussein, F. Degradation of photovoltaic cell performance due to dust deposition on to its surface. Renew. Energy 1993, 3, 585–590. [Google Scholar] [CrossRef]
  8. Figgis, B.; Ennaoui, A.; Ahzi, S.; Rémond, Y. Review of PV soiling particle mechanics in desert environments. Renew. Sustain. Energy Rev. 2017, 76, 872–881. [Google Scholar] [CrossRef]
  9. Figgis, B. Solar Test Facility in Doha, Qatar. 2016. Available online: https://www.researchgate.net/publication/309619145_Solar_Test_Facility_in_Doha_Qatar?channel=doi&linkId=581a3a5608ae30a2c01c9edf&showFulltext=true (accessed on 13 September 2021). [CrossRef]
  10. Bachour, D.; Perez-Astudillo, D. Deriving solar direct normal irradiance using lidar-ceilometer. Sol. Energy 2014, 110, 316–324. [Google Scholar] [CrossRef]
  11. Zheng, C.; Zhao, C.; Zhu, Y.; Wang, Y.; Shi, X.; Wu, X.; Chen, T.; Wu, F.; Qiu, Y. Analysis of influential factors for the relationship between PM2.5 and AOD in Beijing. Atmos. Chem. Phys. 2017, 17, 13473–13489. [Google Scholar] [CrossRef] [Green Version]
  12. Schaap, M.; Apituley, A.; Timmermans, R.; Koelemeijer, R.; de Leeuw, G. Exploring the relation between aerosol optical depth and PM2.5 at Cabauw, the Netherlands. Atmos. Chem. Phys. 2009, 9, 909–925. [Google Scholar] [CrossRef] [Green Version]
  13. Sioris, C.; Abboud, I.; Fioletov, V.; Mclinden, C. AEROCAN, the Canadian sub-network of AERONET: Aerosol monitoring and air quality applications. Atmos. Environ. 2017, 167, 444–457. [Google Scholar] [CrossRef]
  14. Srivastava, N. Association of modeled PM2.5 with aerosol optical depth: Model versus satellite. Nat. Hazards 2020, 102, 689–705. [Google Scholar] [CrossRef]
  15. Martín Pomares, L.; Polo, J.; Perez-Astudillo, D.; Bachour, D. Assessment of Daily Atmospheric Turbidity Databases Using Aerosol Optical Depth and Direct Normal Irradiance Measurements. In Proceedings of the Solar World Congress 2015, Daegu, Republic of Korea, 8–12 November 2015. [Google Scholar] [CrossRef]
  16. Bachour, D.P.-A. Dust impact on shortwave solar radiation: Case study of dust events in Doha, Qatar. In Proceedings of the 6th International Conference Energy & Meteorology, Copenhagen, Denmark, 24–27 June 2019. [Google Scholar]
  17. Benedetti, A.; Morcrette, J.; Boucher, O.; Dethof, A.; Engelen, R.; Fisher, M.; Flentjes, H.; Huneeus, N.; Jones, L.; Kaiser, J.; et al. Aerosol Analysis and Forecast in the ECMWF Integrated Forecast System: Data Assimilation; ECMWF Technical Memoranda: Boston, MA, USA, 2008. [Google Scholar]
  18. Benedetti, A.; Morcrette, J.; Boucher, O.; Dethof, A.; Engelen, R.; Fisher, M.; Flentjes, H.; Huneeus, N.; Jones, L.; Kaiser, J.; et al. Aerosol Analysis and Forecast in the ECMWF Integrated Forecast System: Forward Modelling; ECMWF Technical Memoranda: Boston, MA, USA, 2008. [Google Scholar]
  19. Bachour, D.; Perez-Astudillo, D.; Martín Pomares, L. Study of Soiling on Pyranometers in Desert Conditions. In Proceedings of the EuroSun 2016, Palma de Mallorca, Spain, 11–14 October 2016. [Google Scholar]
  20. Figgis, B.; Ennaoui, A.; Guo, B.; Javed, W.; Chen, E. Outdoor soiling microscope for measuring particle deposition and resuspension. Sol. Energy 2016, 137, 158–164. [Google Scholar] [CrossRef]
  21. Aïssa, B.; Scabbia, G.; Figgis, B.; Garcia-Lopez, J.; Bermudez Benito, V. PV-soiling field-assessment of Mars™ optical sensor operating in the harsh desert environment of the state of Qatar. Sol. Energy 2022, 239, 139–146. [Google Scholar] [CrossRef]
  22. Aïssa, B.; Isaifan, R.; Madhavan, V.E.; Abdallah, A. Structural and physical properties of the dust particles in Qatar and their influence on the PV panel performance. Sci. Rep. 2016, 6, 31467. [Google Scholar] [CrossRef] [Green Version]
  23. Chen, G.; Ziemba, L.D.; Chu, D.A.; Thornhill, K.L.; Schuster, G.L.; Winstead, E.L.; Diskin, G.S.; Ferrare, R.A.; Burton, S.P.; Ismail, S.; et al. Observations of Saharan dust microphysical and optical properties from the Eastern Atlantic during NAMMA airborne field campaign. Atmos. Chem. Phys. Discuss. 2010, 11, 723–740. [Google Scholar] [CrossRef] [Green Version]
  24. Reid, E.A.; Reid, J.S.; Meier, M.M.; Dunlap, M.R.; Cliff, S.S.; Broumas, A.; Perry, K.; Maring, H. Characterization of African dust transported to Puerto Rico by individual particle size segregated bulk analysis. J. Geophys. Res. 2003, 108, 8591. [Google Scholar] [CrossRef]
  25. Yilbas, B.S.; Ali, H.; Khaled, M.M.; Al-Aqeeli, N.; Abu-Dheir, N.; Varanasi, K.K. Influence of dust and mud on the optical, chemical, and mechanical properties of a PV protective glass. Sci. Rep. 2015, 5, 15833. [Google Scholar] [CrossRef] [Green Version]
  26. Coz, E.; Gómez-Moreno, F.; Casuccio, G.; Artíñano, B. Variations on morphology and elemental composition of mineral dust particles from local, regional, and long-range transport meteorological scenarios. J. Geophys. Res. 2010, 115, D12204. [Google Scholar] [CrossRef]
  27. Engelbrecht, J.; Moosmüller, H.; Pincock, S.; Jayanty, R.M.; Lersch, T.; Casuccio, G. Technical Note: Mineralogical, chemical, morphological, and optical interrelationships of mineral dust re-suspensions. Atmos. Chem. Phys. Discuss. 2016, 16, 10809–10830. [Google Scholar] [CrossRef] [Green Version]
  28. Al-Hasan, A.Y. A new correlation for direct beam solar radiation received by photovoltaic panel with sand dust accumulated on its surface. Sol. Energy 1998, 11, 323–333. [Google Scholar] [CrossRef]
  29. Bukowiecki, N.; Lienemann, P.; Hill, M.; Furger, M.; Richard, A.; Amato, F.; Prévôt, A.; Baltensperger, U.; Buchmann, B.; Gehrig, R. PM10 emission factors for non-exhaust particles generated by road traffic in an urban street canyon and along a freeway in Switzerland. Atmos. Environ. 2010, 44, 2330–2340. [Google Scholar] [CrossRef]
  30. Li, T.-C.; Yuan, C.-S.; Huang, H.-C.; Lee, C.-L.; Wu, S.-P.; Tong, C. Inter-comparison of Seasonal Variation, Chemical Characteristics, and Source Identification of Atmospheric Fine Particles on Both Sides of the Taiwan Strait. Sci. Rep. 2016, 6, 22956. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Ma, M.M.; Liu, X.M.; Pillans, B.; Li, P.Y.; Lu, B.; Hu, S.Y. Magnetic properties particle-size analysis of dust-storm samples collected in Lanzhou Sydney. Aust. J. Earth Sci. 2014, 61, 765–774. [Google Scholar] [CrossRef]
  32. Micheli, L.; Muller, M. An investigation of the key parameters for predicting PV soiling losses. Prog. Photovolt. Res. Appl. 2017, 25, 291–307. [Google Scholar] [CrossRef]
  33. Perlwitz, J.; Pérez García-Pando, C.; Miller, R. Predicting the mineral composition of dust aerosols—Part 1: Representing key processes. Atmos. Chem. Phys. 2015, 15, 11593–11627. [Google Scholar] [CrossRef] [Green Version]
  34. Lawrence, C.; Neff, J. The contemporary physical and chemical flux of Aeolian dust: A synthesis of direct measurements of dust deposition. Chem. Geol. 2009, 267, 46–63. [Google Scholar] [CrossRef]
  35. Ilse, K.; Figgis, B.; Naumann, V.; Hagendorf, C.; Bagdahn, J. Fundamentals of soiling processes on photovoltaic modules. Renew. Sustain. Energy Rev. 2018, 98, 239–254. [Google Scholar] [CrossRef]
  36. García, M.; Marroyo, L.; Pigueiras, E.; Perez, M. Soiling and other optical losses in solar-tracking PV plants in Navarra. Prog. Photovolt. Res. Appl. 2011, 19, 211–217. [Google Scholar] [CrossRef] [Green Version]
  37. Rollingwash. Efficient and wWater-Saving Cleaning of the PV Solar Systems with TF-4 Technology I. General Problems of Solar Systems Cleaning. Available online: https://docplayer.net/42360656-Efficient-and-water-saving-cleaning-of-the-pv-solar-systems-with-tf-4-technology.html (accessed on 9 January 2021).
  38. Kimber, A.; Mitchell, L.; Nogradi, S.; Wenger, H. The Effect of Soiling on Large Grid-Connected Photovoltaic Systems in California and the Southwest Region of the United States. In Proceedings of the Conference Record of the 2006 IEEE 4th World Conference on Photovoltaic Energy Conversion, Waikoloa, HI, USA, 7–12 May 2006; Volume 2. [Google Scholar]
  39. Figgis, B.; Guo, B.; Javed, W.; Ilse, K.; Ahzi, S.; Remond, Y. Time-of-day and Exposure Influences on PV Soiling. In Proceedings of the International Renewable and Sustainable Energy Conference, Tangier, Morocco, 4–7 December 2017. [Google Scholar] [CrossRef]
  40. Ilse, K.; Micheli, L.; Figgis, B.W.; Lange, K.; Daßler, D.; Hanifi, H.; Wolfertstetter, F.; Naumann, V.; Hagendorf, C.; Gottschalg, R.; et al. Techno-Economic Assessment of Soiling Losses and Mitigation Strategies for Solar Power Generation. Joule 2019, 3, 2303–2321. [Google Scholar] [CrossRef] [Green Version]
  41. Isaifan, R.; Johnson, D.; Ackermann, L.; Figgis, B.; Ayoub, M. Evaluation of the adhesion forces between dust particles and photovoltaic module surfaces. Sol. Energy Mater. Sol. Cells 2019, 191, 413–421. [Google Scholar] [CrossRef]
  42. Aramrak, S.; Flury, M.; Harsh, J. Detachment of Deposited Colloids by Advancing and Receding Air-Water Interfaces. Langmuir 2011, 27, 9985–9993. [Google Scholar] [CrossRef] [PubMed]
  43. Quan, Y.-Y.; Zhang, L.-Z.; Qi, R.-H.; Cai, R. Self-cleaning of Surfaces: The Role of Surface Wettability and Dust Types. Sci. Rep. 2016, 6, 38239. [Google Scholar] [CrossRef] [Green Version]
  44. Pitois, O.; Chateau, X. Small Particle at a Fluid Interface: Effect of Contact Angle Hysteresis on Force and Work of Detachment. Langmuir 2002, 18, 9751–9756. [Google Scholar] [CrossRef]
  45. Butt, H.-J.; Kappl, M. Normal capillary forces. Adv. Colloid Interface Sci. 2008, 146, 48–60. [Google Scholar] [CrossRef]
  46. Moutinho, H.; Jiang, C.-S.; To, B.; Perkins, C.; Muller, M.; Al-Jassim, M.; Simpson, L. Adhesion mechanisms on solar glass: Effects of relative humidity, surface roughness, and particle shape and size. Sol. Energy Mater. Sol. Cells 2017, 172, 145–153. [Google Scholar] [CrossRef]
  47. Visser, J. Adhesion of colloidal particles. Surf. Colloid Sci. 1976, 8, 3–79. [Google Scholar]
  48. Cleaver, J.; Tyrrell, J. The Influence of Relative Humidity on Particle Adhesion—A Review of Previous Work and the Anomalous Behaviour of Soda-lime Glass. KONA Powder Part. J. 2004, 22, 9–22. [Google Scholar] [CrossRef] [Green Version]
  49. Figgis, B.; Nouviaire, A.; Wubulikasimu, Y.; Javed, W.; Guo, B.; Ait-Mokhtar, A.; Belarbi, R.; Ahzi, S.; Rémond, Y.; Ennaoui, A. Investigation of factors affecting condensation on soiled PV modules. Sol. Energy 2018, 159, 488–500. [Google Scholar] [CrossRef]
  50. Ilse, K.K.; Figgis, B.W.; Werner, M.; Naumann, V.; Hagendorf, C.; Pöllmann, H.; Bagdahn, J. Comprehensive analysis of soiling and cementation processes on PV modules in Qatar. Sol. Energy Mater. Sol. Cells 2018, 186, 309–323. [Google Scholar] [CrossRef]
  51. Mazumder, M.K.; Horenstein, M.N.; Heiling, C.; Stark, J.W.; Sayyah, A.; Yellowhair, J.; Raychowdhury, A. Environmental Degradation of the Optical Surface of PV Modules and Solar Mirrors by Soiling and High RH and Mitigation Methods for Minimizing Energy Yield Losses. In Proceedings of the IEEE 42nd Photovoltaic Specialist Conference (PVSC), New Orleans, LA, USA, 14–19 June 2015. [Google Scholar] [CrossRef]
  52. Javed, W.; Guo, B.; Figgis, B. Modeling of photovoltaic soiling loss as a function of environmental variables. Sol. Energy 2017, 157, 397–407. [Google Scholar] [CrossRef]
  53. Heck, M.; Köhl, M. Simulation of Water Vapor Ingress into PV-Modules under Different Climatic Conditions. J. Mater. 2013, 2013, 1–7. [Google Scholar]
  54. Lombardo, T.; Ionescu, A.; Chabas, A.; Lefèvre, R.A.; Ausset, P.; Candau, Y. Dose–response function for the soiling of silica–soda–lime glass due to dry deposition. Sci. Total Environ. 2009, 408, 976–984. [Google Scholar] [CrossRef] [PubMed]
  55. Ilse, K.; Khan, M.; Naumann, V.; Hagendorf, C. Dew as a Detrimental Influencing Factor for Soiling of PV Modules. IEEE J. Photovolt. 2018, 9, 287–294. [Google Scholar] [CrossRef]
  56. Cuddihy, E. Theoretical considerations of soil retention. Sol. Energy Mater. 1980, 3, 21–33. [Google Scholar] [CrossRef]
  57. Mani, M.; Pillai, R. Impact of dust on solar photovoltaic (PV) performance: Research status, challenges and recommendations. Renew. Sustain. Energy Rev. 2010, 14, 3124–3131. [Google Scholar] [CrossRef]
  58. 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.; Hanriot, S.D.M.; Cruz, L.R.D.O. Fundamental studies of the adhesion of dust to PV module chemical and physical relationships at the microscale. IEEE J. Photovolt. 2015, 6, 719–729. [Google Scholar] [CrossRef]
  59. Ilse, K.; Werner, M.; Naumann, V.; Figgis, B.W.; Hagendorf, C.; Bagdahn, J. Microstructural analysis of the cementation process during soiling on glass surfaces in arid and semi-arid climates. Phys. Status Solidi-Rapid Res. Lett. 2016, 10, 525–529. [Google Scholar] [CrossRef]
  60. Perkins, C.L.; Muller, M.; Simpson, L. Laboratory Studies of Particle Cementation and PV module Soiling. In Proceedings of the 2017 IEEE 44th Photovoltaic Specialist Conference (PVSC), Washington, DC, USA, 25–30 June 2017; pp. 2294–2297. [Google Scholar] [CrossRef]
  61. Figgis, B.; Goossens, D.; Guo, B.; Ilse, K. Effect of tilt angle on soiling in perpendicular wind. Sol. Energy 2019, 194, 294–301. [Google Scholar] [CrossRef]
  62. Smith, K.; Goossens, D. Wind tunnel simulations of aeolian dust deposition on thermic solar collectors. Appl. Sol. Energy (Engl. Transl. Geliotekhnika) 1995, 31, 75–89. [Google Scholar]
  63. Goossens, D. Soiling of Photovoltaic Modules: Size Characterization of the Accumulated Dust. J. Clean Energy Technol. 2019, 7, 25–31. [Google Scholar] [CrossRef]
  64. Jaszczur, M.; Koshti, A.; Nawrot, W.; Sędor, P. An investigation of the dust accumulation on photovoltaic panels. Environ. Sci. Pollut. Res. 2020, 27, 2001–2014. [Google Scholar] [CrossRef]
  65. Javed, W.; Yiming, W.; Figgis, B.; Guo, B. Characterization of dust accumulated on photovoltaic panels in Doha, Qatar. Sol. Energy 2017, 142, 123–135. [Google Scholar] [CrossRef]
  66. Mastekbayeva, G.; Kumar, S. Effect of dust on the transmittance of low-density polyethylene glazing in a tropical climate. Sol. Energy 2000, 68, 135–141. [Google Scholar] [CrossRef]
  67. Elminir, H.K.; Ghitas, A.E.; Hamid, R.; El-Hussainy, F.; Beheary, M.; Abdel-Moneim, K.M. Effect of dust on the transparent cover of solar collectors. Energy Convers. Manag. 2006, 47, 3192–3203. [Google Scholar] [CrossRef]
  68. Hegazy, A. Effect of dust accumulation on solar transmittance through glass covers of plate-type collectors. Renew. Energy 2001, 22, 525–540. [Google Scholar] [CrossRef]
  69. Alnaser, N.W.; Dakhel, A.A.; Al Othman, M.J.; Batarseh, I.; Lee, J.K.; Najmaii, S.; Alnaser, W.E. Dust Accumulation Study on the Bapco 05 MW p PVProject at University of Bahrain. Int. J. Power Energy Syst. 2015, 2, 35–54. [Google Scholar]
  70. Tanesab, J.; Parlevliet, D.; Whale, J.; Urmee, T. The effect of dust with different morphologies on the performance degradation of photovoltaic modules. Sustain. Energy Technol. Assess. 2019, 31, 347–354. [Google Scholar] [CrossRef]
  71. Javed, W.; Guo, B.; Figgis, B.; Aïssa, B. Dust potency in the context of solar photovoltaic (PV) soiling loss. Sol. Energy 2021, 220, 1040–1052. [Google Scholar] [CrossRef]
  72. Besson, P.; Munoz, C.; Ramirez-Sagner, G.; Salgado, M.; Escobar, R.; Platzer, W. Long-Term Soiling Analysis for Three Photovoltaic Technologies in Santiago Region. IEEE J. Photovolt. 2017, 7, 1755–1760. [Google Scholar] [CrossRef]
  73. Javed, W.; Guo, B.; Figgis, B.; Pomares, L.M.; Aïssa, B. Multi-year field assessment of seasonal variability of photovoltaic soiling and environmental factors in a desert environment. Sol. Energy 2020, 211, 1392–1402. [Google Scholar] [CrossRef]
  74. Mondal, S.; Mondal, A.K.; Sharma, A.; Devalla, V.; Rana, S.; Kumar, S.; Pandey, J.K. An Overview of Cleaning and Prevention Processes for Enhancing Efficiency of Solar Photovoltaic Panels. Curr. Sci. 2018, 115, 1065–1077. [Google Scholar] [CrossRef]
  75. Einhorn, A.; Micheli, L.; Miller, D.C.; Simpson, L.J.; Moutinho, H.R.; To, B.; Lanaghan, C.L.; Muller, M.T.; Toth, S.; John, J.J.; et al. Evaluation of Soiling Potential Mitigation Approaches on Photovoltaic Glass. IEEE J. Photovolt. 2018, 9, 233–239. [Google Scholar] [CrossRef]
  76. Gallagher, B. Rise of the Machines: Solar Module-Washing Robots. 2018. Available online: https://www.woodmac.com/our-expertise/focus/Power--Renewables/rise-of-the-machines-solar-module-washing-robots/ (accessed on 13 September 2021).
  77. Maghami, M.R.; Hizam, H.; Gomes, C.; Radzi, M.A.; Rezadad, M.I.; Hajighorbani, S. Power loss due to soiling on solar panel: A review. Renew. Sustain. Energy Rev. 2016, 59, 1307–1316. [Google Scholar]
  78. Alghamdi, A.S.; Bahaj, A.S.; Blunden, L.S.; Wu, Y. Dust Removal from Solar PV Modules by Automated Cleaning Systems. Energies 2019, 12, 2923. [Google Scholar] [CrossRef] [Green Version]
  79. Hossain, M.I.; Ali, A.; Bermudez Benito, V.; Figgis, B.; Aïssa, B. Anti-Soiling Coatings for Enhancement of PV Panel Performance in Desert Environment: A Critical Review and Market Overview. Materials 2022, 15, 7139. [Google Scholar] [CrossRef]
  80. Costa, S.; Cardoso Diniz, A.S.; Kazmerski, L. Solar energy dust and soiling R&D progress: Literature review update for 2016. Renew. Sustain. Energy Rev. 2017, 82, 2504–2536. [Google Scholar]
  81. Costa, S.; Cardoso Diniz, A.S.; Kazmerski, L. Dust and soiling issues and impacts relating to solar energy systems: Literature review update for 2012–2015. Renew. Sustain. Energy Rev. 2016, 63, 33–61. [Google Scholar] [CrossRef] [Green Version]
  82. Alkandary, A.T.; Al-Asfour, F.G.A. Comparative Assessment of Anti-soiling Nanocoated Photovoltaic Modules in an Arid Desert Environment. In Proceedings of the 35th European Photovoltaic Solar Energy Conference and Exhibition, Brussels, Belgium, 24–28 September 2018. [Google Scholar]
  83. Brito, P.P.; Diniz, A.S.; Kazmerski, L.L. Materials design and discovery: Potential for application to soiling mitigation in photovoltaic systems. Sol. Energy 2019, 183, 791–804. [Google Scholar] [CrossRef]
  84. Abrams, Z.; Gonsalves, P.; Brophy, B.; Posbic, J. Field and Lab Verification of Hydrophobic Anti-Reflective and Anti-Soiling Coatings on Photovoltaic Glass. In Proceedings of the 29th European Photovoltaic Solar Energy Conference and Exhibition, Amsterdam, The Netherlands, 22–26 September 2014; pp. 2759–2764. [Google Scholar] [CrossRef]
  85. Bahattab, M.A.; Alhomoudi, I.A.; Alhussaini, M.I.; Mirza, M.; Hegmann, J.; Glaubitt, W.; Löbmann, P. Anti-soiling surfaces for PV applications prepared by sol-gel processing: Comparison of laboratory testing and outdoor exposure. Sol. Energy Mater. Sol. Cells 2016, 157, 422–428. [Google Scholar] [CrossRef]
  86. Brophy, B.; Abrams, Z.; Gonsalves, P.; Christy, K. Field Performance and Persistance of Anti-Soiling Coatings on Photovoltaic Glass. In Proceedings of the 31st European Photovoltaic Solar Energy Conference and Exhibition, Hamburg, Germany, 14–18 September 2015. [Google Scholar]
  87. Ilse, K.; Khan, M.Z.; Voicu, N.; Naumann, V.; Hagendorf, C.; Bagdahn, J. Advanced performance testing of anti-soiling coatings—Part II: Particle-size dependent analysis for physical understanding of dust removal processes and determination of adhesion forces. Sol. Energy Mater. Sol. Cells 2019, 202, 110049. [Google Scholar] [CrossRef]
  88. Ferretti, N.; Ilse, K.; Sönmez, A.; Hagendorf, C.; Berghold, J. Investigation on the Impact of Module Cleaning on the Antireflection Coating. In Proceedings of the 32nd European Photovoltaic Solar Energy Conference and Exhibition, Munich, Germany, 20–24 June 2016. [Google Scholar] [CrossRef]
  89. Wilke, K.; Preston, D.; Lu, Z.; Wang, E. Toward Condensation-Resistant Omniphobic Surfaces. ACS Nano 2018, 12, 11013–11021. [Google Scholar] [CrossRef] [PubMed]
  90. Isbilir, K.; Lisco, F.; Womack, G.; Abbas, A.; Walls, J. Testing of an Anti-Soiling Coating for PV Module Cover Glass. In Proceedings of the 2018 IEEE 7th World Conference on Photovoltaic Energy Conversion (WCPEC) (A Joint Conference of 45th IEEE PVSC, 28th PVSEC & 34th EU PVSEC, Waikoloa, HI, USA, 10–15 June 2018. [Google Scholar] [CrossRef] [Green Version]
  91. Toth, S.; Muller, M.; Miller, D.C.; Moutinho, H.; To, B.; Micheli, L.; Linger, J.; Engtrakul, C.; Einhorn, A.; Simpson, L. Soiling and cleaning: Initial observations from 5-year photovoltaic glass coating durability study. Sol. Energy Mater. Sol. Cells 2018, 185, 375–384. [Google Scholar] [CrossRef]
  92. Ilse, K.K.; Rabanal, J.; Schonleber, L.; Khan, M.Z.; Naumann, V.; Hagendorf, C.; Bagdahn, J. Comparing Indoor Outdoor Soiling Experiments for Different Glass Coatings Microstructural Analysis of Particle Caking Processes. IEEE J. Photovolt. 2018, 8, 203–209. [Google Scholar] [CrossRef]
  93. Steensma, G.; Román, R.; Marshall, C.; Bermejo, J.; Iyer, K.; Al-Hajraf, S.; Al-Qattan, A. Shagaya renewable energy park project. In Proceedings of the 24th SolarPACES International Conference, Casablanca, Morocco, 2–5 October 2018; Volume 2126. [Google Scholar]
  94. PV Cleaning Solution Company NOMADD Desert Solar Solutions Secures Funding Round from Saudi Arabian Investor CEPCO. 2018. Available online: http://www.nomaddesertsolar.com/news (accessed on 15 October 2020).
  95. Albadwawi, O.; John, J.; Elhassan, Y.; Albanna, F.; Almamari, F.; Alqassim, A.; Alnuaimi, A. Quantification of spectral losses of Natural soiling and detailed microstructural analysis of Dust collected from Different locations in Dubai, UAE. In Proceedings of the 2019 IEEE 46th Photovoltaic Specialists Conference (PVSC), Chicago, IL, USA, 16–21 June 2019. [Google Scholar] [CrossRef]
  96. Figgis, B. QEERI Solar Test Facility 5-Year Report; Qatar Environment and Energy Research Institute: Ar-Rayyan, Qatar, 2018. [Google Scholar] [CrossRef]
  97. Nitz, P. Investigations into Soiling and Degradation. Available online: https://www.ise.fraunhofer.de/en/business-areas/photovoltaics/iii-v-and-concentrator-photovoltaics/concentrator-optics/investigations-into-soiling-and-degradation.html (accessed on 3 March 2021).
  98. Simpson, L.J.; Muller, M.; Deceglie, M.; Moutinho, H.; Perkins, C.; Jiang, C.S.; Miller, D.C.; Micheli, L.; Tamizhmani, G.; Tatapudi, S.R.V.; et al. NREL Efforts to Address Soiling on PV Modules. In Proceedings of the 2017 IEEE 44th Photovoltaic Specialist Conference (PVSC), Washington, DC, USA, 25–30 June 2017; pp. 2789–2793. [Google Scholar] [CrossRef] [Green Version]
  99. Cabrera, E.; Araya, F.; Schneider, A.; Marzo, A.; Roescu, R.; Rabanal, J.; Ferrada, P.; Ayala, P.; Berrian, D.; Libal, J.; et al. AtaMoS-TeC Project: The bifacial institute for desert PV. In Proceedings of the Bifi PV Workshop, Denver, CO, USA, 10–11 September 2018. [Google Scholar]
  100. King, B. Overview of Sandia’s Soiling Program: Experimental Methods and Framework for a Quantitative Soiling Model. Sandia National Laboratories Internal Report, Published on 22 October 2015. Available online: https://www.osti.gov/servlets/purl/1331410 (accessed on 2 February 2020).
  101. Welcome to PVReliability.org Our Collaborative R&D Program Aimed at Ensuring the Reliability of Solar PV Systems. Available online: http://www.pvreliability.org/ (accessed on 4 March 2021).
  102. Institute of Solar Research in the EU-Funded Project WASCOP: Soiling Prediction Model, Tests of mirror Coatings and Innovative Cleaning Methods for Solar Mirrors. 2016. Available online: https://www.dlr.de/sf/en/desktopdefault.aspx/tabid-10436/20174_read-47350/ (accessed on 4 March 2021).
  103. Herrmann, W. Impact of Soiling on PV Module Performance for Various Climates. In Proceedings of the 4th PV Performance Modelling and Monitoring Workshop, Cologne, Germany, 22–23 October 2015. [Google Scholar]
  104. Soiling: Task Group 12. Available online: https://www.pvqat.org/project-status/task-group-12.html (accessed on 4 March 2021).
  105. Available online: https://cost-indust.eu/the-action/working-groups (accessed on 6 April 2021).
  106. Empowering Solar World Leader in Robotic Solar Cleaning. Available online: https://www.ecoppia.com/ (accessed on 6 April 2021).
  107. Automatic and Semiautomatic Portable Washing Systems for PV Plants. Available online: http://www.washpanel.com/en/ (accessed on 6 April 2021).
  108. ERICWESOFF; HT SunPower Cleans up Solar with Acquisition of Greenbotics. 2013. Available online: https://www.greentechmedia.com/articles/read/sunpower-cleans-up-solar-with-acquisiton-of-greenbotics (accessed on 6 April 2021).
  109. First Solar Rooftop in the UAE with 100% Robotic Maintenance Is Completed. 2019. Available online: http://www.mesia.com/2019/02/18/first-solar-rooftop-in-the-uae-with-100-robotic-maintenance-is-completed/ (accessed on 29 January 2021).
  110. Available online: https://www.serbot.ch/en/ (accessed on 29 January 2021).
  111. Solar Panel Automatic Cleaning Robot with Self-Propelling Guide Rail (the World’s First). 2014. Available online: https://www.skyrobot.co.jp/en/solrider.html (accessed on 28 January 2021).
  112. Solar Hybrid Technology. Available online: https://www.enerwhere.com/technology (accessed on 28 January 2021).
  113. BladeRanger Autonomous Cleaning and Inspection of Solar Power Installations: Let the Sunshine In! Available online: https://bladeranger.com/ (accessed on 11 January 2021).
  114. Wolfertstetter, F.; Fonk, R.; Prahl, C.; Röger, M.; Wilbert, S.; Fernández-Reche, J. Airborne soiling measurement of entire solar fields with Qfly. In Proceedings of the International Conference on Concentrating Solar Power and Chemical Energy Systems, Daegu, South Korea, 1–4 October 2019. [Google Scholar]
  115. SMP100 Solar-Module Performance Monitoring System. Available online: https://www.campbellsci.com/smp100 (accessed on 3 May 2021).
  116. DustIQ Soiling Monitoring System. Available online: https://www.kippzonen.com/Product/419/DustIQ-Soiling-Monitoring-System (accessed on 9 January 2022).
  117. Available online: https://norcalcontrols.net/ (accessed on 9 January 2022).
  118. Mars Optical Soiling Sensor—Revolutionary New Product. Available online: http://www.atonometrics.com/mars-optical-soiling-sensor-revolutionary-new-product/ (accessed on 9 January 2022).
  119. Soiling Measurement KIT. Available online: https://www.kintech-engineering.com/catalogue/soiling-measurement-kit/soiling-measurement-kit/ (accessed on 9 January 2022).
  120. Quantifying Solar O&M’S Dirty Side. Available online: https://europe.solar-asset.management/new-updates-source/2017/10/19/quantifying-solar-oms-dirty-side (accessed on 9 January 2022).
  121. Solar Energy 3.0. Available online: https://www.innoenergy.com/discover-innovative-solutions/sustainable-products-services/electricity-heat-production/solar-pv/solar-energy-30/ (accessed on 9 January 2022).
  122. Available online: https://trackso.in/ (accessed on 9 January 2022).
  123. CHRIS COWELL. Anti-Soiling Coating That Can Increase Solar Panel Output Heads to Commercial Deployment. 2019. Available online: https://solarbuildermag.com/products/anti-soiling-coating-that-can-increase-solar-panel-output-heads-to-commercial-deployment/ (accessed on 1 February 2022).
  124. Anti-Soiling Coatings. Available online: https://www.dsm.com/dsm-in-solar/en_US/technologies/anti-soiling-coatings-for-solar-glass.html (accessed on 1 February 2022).
  125. Nano-Coatings for Water Control Moisture·Condensation·Wetting. Available online: https://lotusleafcoatings.com/ (accessed on 10 June 2021).
  126. Figgis, B.; Ilse, K. Anti-Soiling Potential of 1-Axis PV Trackers. 2019. Available online: https://www.eupvsec-proceedings.com/proceedings?paper=48143 (accessed on 6 February 2021). [CrossRef]
  127. Gu, W.; Ma, T.; Ahmed, S.; Zhang, Y.; Peng, J. A comprehensive review and outlook of bifacial photovoltaic (bPV) technology. Energy Convers. Manag. 2020, 223, 113283. [Google Scholar] [CrossRef]
  128. Guo, S. Vertically mounted bifacial photovoltaic modules: A global analysis. Energy 2013, 61, 447–454. [Google Scholar] [CrossRef]
  129. Ito, M.; Gerritsen, E. Geographical Mapping of the Performance of Vertically Installed Bifacial Modules. In Proceedings of the 32nd European Photovoltaic Solar Energy Conference and Exhibition, Munich, Germany, 20–24 June 2016. [Google Scholar]
  130. Kopecek, R.; Libal, J. Towards large-scale deployment of bifacial photovoltaics. Nat. Energy 2018, 3, 443–446. [Google Scholar] [CrossRef]
  131. Ullah, A.; Amin, A.; Kazmi, T.; Saleem, M.; Butt, N. Investigation of soiling effects, dust chemistry and optimum cleaning schedule for PV modules in Lahore, Pakistan. Renew. Energy 2020, 150, 456–468. [Google Scholar] [CrossRef]
  132. Rabanal Arabach, J.; Schneider, A.; Mrcarica, M.; Kopecek, R.; Heckmann, M. The Need of Frameless Mounting Structures for Vertical Mounting of Bifacial PV Modules. In Proceedings of the 32nd EU PVSEC, Munich, Germany, 20–24 June 2016. [Google Scholar] [CrossRef]
  133. Bhaduri, S.; Kottantharayil, A. Mitigation of Soiling by Vertical Mounting of Bifacial Modules. IEEE J. Photovolt. 2018, 9, 240–244. [Google Scholar] [CrossRef]
  134. Luque, E.; Antonanzas, F.; Escobar, R. Effect of soiling in bifacial PV modules and cleaning schedule optimization. Energy Convers. Manag. 2018, 174, 615–625. [Google Scholar] [CrossRef]
  135. Dassler, D.; Malik, S.; Figgis, B.; Bagdahn, J.; Ebert, M. Bifacial gain simulations of modules and systems under desert conditions. In Proceedings of the 4th bifi PV Workshop, Konstanz, Germany, 25–26 October 2017. [Google Scholar]
  136. Rodríguez-Gallegos, C.D.; Liu, H.; Gandhi, O.; Singh, J.P.; Krishnamurthy, V.; Kumar, A.; Stein, J.S.; Wang, S.; Li, L.; Reindl, T.; et al. Global Techno-Economic Performance of Bifacial and Tracking Photovoltaic Systems. Joule 2020, 4, 1514–1541. [Google Scholar] [CrossRef]
  137. Cabrera, E.; Araya, F.; Schneider, A.; Marzo, A.; Roescu, R.; Rabanal, J.; Ferrada, P.; Ayala, P.; Berrian, D.; Libal, J.; et al. AtaMoS TeC Project: Soiling Impact on Bifacial Modules with Different Mounting Geometry in the Atacama Desert in Chile. In Proceedings of the 2018 IEEE 7th World Conference on Photovoltaic Energy Conversion (WCPEC) (A Joint Conference of 45th IEEE PVSC, 28th PVSEC & 34th EU PVSEC), Waikoloa, HI, USA, 10–15 June 2018. [Google Scholar] [CrossRef]
  138. Moehlecke, A.; Febras, F.; Zanesco, I. Electrical performance analysis of PV modules with bifacial silicon solar cells and white diffuse reflector. Sol. Energy 2013, 96, 253–262. [Google Scholar] [CrossRef]
  139. Schmela, M.; Beauvais, A.; Chevillard, N.; Paredes, M.; Heisz, M.; Rossi, R. Global Market Outlook For Solar Power. Glob. Mark. Outlook 2018, 1–92. Available online: https://www.photovoltaique.info/media/filer_public/2d/a7/2da79273-50e7-4099-9fe8-c1c3e4a47590/solarpower-europe-global-market-outlook-2019-2023.pdf (accessed on 6 February 2021).
  140. Jäger-Waldau, A. JRC Science for Policy Report; PV Status Report; European Commission: Brussels, Belgium, 2017; pp. 3–83. [Google Scholar] [CrossRef]
  141. Jäger-Waldau, A. Snapshot of photovoltaics-February 2018. EPJ Photovolt. 2018, 9, 6. [Google Scholar] [CrossRef]
  142. Apostoleris, H.; Sgouridis, S.; Stefancich, M.; Chiesa, M. Evaluating the factors that led to low-priced solar electricity projects in the Middle East. Nat. Energy 2018, 3, 1109–1114. [Google Scholar] [CrossRef]
  143. Labordena, M.; Neubauer, D.; Folini, D.; Patt, A.; Lilliestam, J. Blue skies over China: The effect of pollution-control on solar power generation and revenues. PLoS ONE 2018, 13, e0207028. [Google Scholar] [CrossRef]
  144. Boys, B.L.; Martin, R.V.; van Donkelaar, A.; MacDonell, R.J.; Hsu, N.C.; Cooper, M.J.; Yantosca, R.M.; Lu, Z.; Streets, D.G.; Zhang, Q.; et al. Fifteen-Year Global Time Series of Satellite-Derived Fine Particulate Matter. Environ. Sci. Technol. 2014, 48, 11109–11118. [Google Scholar] [CrossRef]
  145. Yang, X.; Jiang, L.; Zhao, W.; Xiong, Q.; Zhao, W.; Yan, X. Comparison of Ground-Based PM2.5 and PM10 Concentrations in China, India, and the U.S. Int. J. Environ. Res. Public Health 2018, 15, 1382. [Google Scholar] [CrossRef] [Green Version]
  146. Butt, E.W.; Turnock, S.T.; Rigby, R.; Reddington, C.L.; Yoshioka, M.; Johnson, J.S.; A Regayre, L.; Pringle, K.J.; Mann, G.W.; Spracklen, D.V. Global and regional trends in particulate air pollution and attributable health burden over the past 50 years. Environ. Res. Lett. 2017, 12, 104017. [Google Scholar] [CrossRef]
  147. Weagle, C.L.; Snider, G.; Li, C.; Van Donkelaar, A.; Philip, S.; Bissonnette, P.; Burke, J.; Jackson, J.; Latimer, R.; Stone, E.; et al. Global Sources of Fine Particulate Matter: Interpretation of PM 2.5 Chemical Composition Observed by SPARTAN using a Global Chemical Transport Model. Environ. Sci. Technol. 2018, 52, 11670–11681. [Google Scholar] [CrossRef]
  148. Al-Kouz, W.; Al-Dahidi, S.; Hammad, B.; Al-Abed, M. Modeling and Analysis Framework for Investigating the Impact of Dust and Temperature on PV Systems’ Performance and Optimum Cleaning Frequency. Appl. Sci. 2019, 9, 1397. [Google Scholar] [CrossRef] [Green Version]
  149. Lopez-Lorente, J.; Polo, J.; Martín-Chivelet, N.; Norton, M.; Livera, A.; Makrides, G.; Georghiou, G.E. Characterizing soiling losses for photovoltaic systems in dry climates: A case study in Cyprus. Sol. Energy 2023, 255, 243. [Google Scholar] [CrossRef]
  150. Micheli, L.; Smestad, G.P.; Bessa, J.G.; Muller, M.; Fernandez, E.F.; Almonacid, F. Tracking soiling losses: Assessment, Uncertainty, and challenges in mapping. IEEE J. Photovolt. 2022, 12, 114. [Google Scholar] [CrossRef]
  151. Bhaduri, S.; Bajhal, R.; Farkade, M.; Mallick, S.; Shiradkar, N.; Kottantharayil, A. Understanding Multiple Stressors Which Degrade Antisoiling Coatings: Combined Effect of Rain, Abrasion, and UV Radiation. IEEE J. Photovoltaics. 2023, 1–7. [Google Scholar] [CrossRef]
  152. IRENA. Renewable Energy Marbifacialket Analysis: GCC 2019; IRENA: Masdar City, Abu Dhabi, 2019. [Google Scholar]
Figure 1. Number of refereed publications on PV soiling per year, along with their associated distribution of the used keywords. Data were collected from January 1999 to November 2022, mainly from Scopus®.
Figure 1. Number of refereed publications on PV soiling per year, along with their associated distribution of the used keywords. Data were collected from January 1999 to November 2022, mainly from Scopus®.
Energies 16 05224 g001
Figure 2. (a) Photo of the QEERI Outdoor Test Facility’s area. (b) Cleaned vs. soiled PV modules surface after 30 days of exposure to outdoor conditions.
Figure 2. (a) Photo of the QEERI Outdoor Test Facility’s area. (b) Cleaned vs. soiled PV modules surface after 30 days of exposure to outdoor conditions.
Energies 16 05224 g002
Figure 3. Soiling degree of a pyranometer operating in Doha, Qatar, showing the decrease of the measured global horizontal radiation per day due to sensor soiling.
Figure 3. Soiling degree of a pyranometer operating in Doha, Qatar, showing the decrease of the measured global horizontal radiation per day due to sensor soiling.
Energies 16 05224 g003
Figure 4. (a) Outdoor soiling microscope with a glass slide serving as a dust collection surface. (b) Lab microscope images of the same soiled surface (left) and the associated outdoor soiling microscope image (right). (c) Condensation droplets image was taken by the outdoor soiling microscope. (d) Removal rate vs. relative humidity and wind speed. Adapted from ref. [20].
Figure 4. (a) Outdoor soiling microscope with a glass slide serving as a dust collection surface. (b) Lab microscope images of the same soiled surface (left) and the associated outdoor soiling microscope image (right). (c) Condensation droplets image was taken by the outdoor soiling microscope. (d) Removal rate vs. relative humidity and wind speed. Adapted from ref. [20].
Energies 16 05224 g004
Figure 5. (a) Mars™ soiling sensor and (b) Dust IQ sensor at PV test field along with clean/soiled reference cell pair. (c) Summary of the daily soiling rate (SR) measured from the data of the Mars™ during different periods. Results were compared to the PV arrays’ daily power losses obtained from two different arrays. Adapted from ref. [21].
Figure 5. (a) Mars™ soiling sensor and (b) Dust IQ sensor at PV test field along with clean/soiled reference cell pair. (c) Summary of the daily soiling rate (SR) measured from the data of the Mars™ during different periods. Results were compared to the PV arrays’ daily power losses obtained from two different arrays. Adapted from ref. [21].
Energies 16 05224 g005
Figure 6. (a) Summary of the calculated SR %, and (b) mismatch (ΔSR) between the data analyzed from the MarsTM sensor and PV arrays. Inset in (b) is the Qatar map showing the two sites where the MarsTM and/or Dust IQ sensors were operated. Adapted from Ref. [21].
Figure 6. (a) Summary of the calculated SR %, and (b) mismatch (ΔSR) between the data analyzed from the MarsTM sensor and PV arrays. Inset in (b) is the Qatar map showing the two sites where the MarsTM and/or Dust IQ sensors were operated. Adapted from Ref. [21].
Energies 16 05224 g006
Figure 7. (a) Typical SEM micrograph of the desert dust particles collected from the OTF PV panels. An example of a particle-size-distribution histogram is shown in the inset. (b) Energy dispersive X-rays spectra, and (c) the associated chemical composition of the dust particles. (d) Representative X-ray diffraction pattern of dust particles. Adapted from ref. [22].
Figure 7. (a) Typical SEM micrograph of the desert dust particles collected from the OTF PV panels. An example of a particle-size-distribution histogram is shown in the inset. (b) Energy dispersive X-rays spectra, and (c) the associated chemical composition of the dust particles. (d) Representative X-ray diffraction pattern of dust particles. Adapted from ref. [22].
Energies 16 05224 g007
Figure 8. Schematic of site adaption highlighting the effect of urban traffic on soiling.
Figure 8. Schematic of site adaption highlighting the effect of urban traffic on soiling.
Energies 16 05224 g008
Figure 9. Various chemical mineral compositions based on XRD analysis of dust particles collected from the ground of various regions over the world [27,28]. Statistics in this figure are adapted from [35].
Figure 9. Various chemical mineral compositions based on XRD analysis of dust particles collected from the ground of various regions over the world [27,28]. Statistics in this figure are adapted from [35].
Energies 16 05224 g009
Figure 10. Influencing factors of the PV soiling: correlational system between the size scale in meters and the time scale in seconds, showing the complexity of the PV soiling phenomenon and its high degrees of freedom. Adapted from ref. [35].
Figure 10. Influencing factors of the PV soiling: correlational system between the size scale in meters and the time scale in seconds, showing the complexity of the PV soiling phenomenon and its high degrees of freedom. Adapted from ref. [35].
Energies 16 05224 g010
Figure 11. Schematic Illustration of important soiling mechanisms. Adapted from reference [40].
Figure 11. Schematic Illustration of important soiling mechanisms. Adapted from reference [40].
Energies 16 05224 g011
Figure 12. Representative scheme of the (a) capillary, (b) van der Waals, (c) Electrostatic, and (d) Gravitational forces between a dust particle and a glass substrate.
Figure 12. Representative scheme of the (a) capillary, (b) van der Waals, (c) Electrostatic, and (d) Gravitational forces between a dust particle and a glass substrate.
Energies 16 05224 g012
Figure 13. (a) Photos taken early in the morning of water condensation formed at the front glass of the PV array. (b) Water condensation on the backside of a PV module installed at QEERI/OTF. Adapted from [49].
Figure 13. (a) Photos taken early in the morning of water condensation formed at the front glass of the PV array. (b) Water condensation on the backside of a PV module installed at QEERI/OTF. Adapted from [49].
Energies 16 05224 g013
Figure 14. Optical OSM photos of various droplets formed on glass and PTFE surfaces. Ambient conditions and 4 h during nighttime conditions were used. Adapted from Ref. [49].
Figure 14. Optical OSM photos of various droplets formed on glass and PTFE surfaces. Ambient conditions and 4 h during nighttime conditions were used. Adapted from Ref. [49].
Energies 16 05224 g014
Figure 15. (a) Dust accumulation on a PV array following a sandstorm in OTF, Doha. (b) Schematic of various surface tilt angles.
Figure 15. (a) Dust accumulation on a PV array following a sandstorm in OTF, Doha. (b) Schematic of various surface tilt angles.
Energies 16 05224 g015
Figure 16. (a) Setup of wind-tunnel tests. (b) Field soiling wind performed in OTF: the left coupon is at 0° tilt, and the right coupon is at −22° facing downwind. (c) Wind-tunnel imaging: structure of soiling layer along the glass coupon in cm (WS = 2 m·s−1, tilt angle = 30°). Adapted from ref. [61].
Figure 16. (a) Setup of wind-tunnel tests. (b) Field soiling wind performed in OTF: the left coupon is at 0° tilt, and the right coupon is at −22° facing downwind. (c) Wind-tunnel imaging: structure of soiling layer along the glass coupon in cm (WS = 2 m·s−1, tilt angle = 30°). Adapted from ref. [61].
Energies 16 05224 g016
Figure 17. Variation of monthly average DAR, RH%, PM10, and WS from January 2015 to October 2016. Adapted from ref. [71].
Figure 17. Variation of monthly average DAR, RH%, PM10, and WS from January 2015 to October 2016. Adapted from ref. [71].
Energies 16 05224 g017
Figure 18. Daily ΔCI % as a function of the number of days. Adapted from ref. [73].
Figure 18. Daily ΔCI % as a function of the number of days. Adapted from ref. [73].
Energies 16 05224 g018
Figure 19. PV cleanness index (CI) of the soiled PV array as a function of time (bi-monthly cleaned modules) from 2014 to 2019. Adapted from ref. [73].
Figure 19. PV cleanness index (CI) of the soiled PV array as a function of time (bi-monthly cleaned modules) from 2014 to 2019. Adapted from ref. [73].
Energies 16 05224 g019
Figure 20. Soiling Mitigation Technologies (a) manual cleaning (b) automated cleaning (c) dew mitigation for cementation control using heating (d) anti-soling coating and (e) electrodynamic screen or shields for repelling dust particles.
Figure 20. Soiling Mitigation Technologies (a) manual cleaning (b) automated cleaning (c) dew mitigation for cementation control using heating (d) anti-soling coating and (e) electrodynamic screen or shields for repelling dust particles.
Energies 16 05224 g020
Figure 21. Photos examples showing the main cleaning technologies, namely manual (left), semi-automatic (middle, including truck-mounted portable robot models), and (right) fully automatic. Reproduced from [40].
Figure 21. Photos examples showing the main cleaning technologies, namely manual (left), semi-automatic (middle, including truck-mounted portable robot models), and (right) fully automatic. Reproduced from [40].
Energies 16 05224 g021
Figure 22. (a) 10 cm × 10 cm glass coupons were mounted on two 360° rotatable 1-axis trackers (left and middle of picture) and one fixed-tilt stand (right). The axes were aligned East-West; as pictured all coupons are tilted South (b) Two trackers with full-size modules were built, to validate results obtained from 10 cm coupons (c) Flux rate of dust to (positive values) or from (negative values) a horizontal glass coupon vs. time of day in Doha, Qatar. Adapted from ref. [126].
Figure 22. (a) 10 cm × 10 cm glass coupons were mounted on two 360° rotatable 1-axis trackers (left and middle of picture) and one fixed-tilt stand (right). The axes were aligned East-West; as pictured all coupons are tilted South (b) Two trackers with full-size modules were built, to validate results obtained from 10 cm coupons (c) Flux rate of dust to (positive values) or from (negative values) a horizontal glass coupon vs. time of day in Doha, Qatar. Adapted from ref. [126].
Energies 16 05224 g022
Figure 23. (a) Soiling loss experiments for three configurations: vertical bifacial (VB), latitude tilt bifacial (LB), and latitude tilt monofacial (LM) modules [127], (b) Monofacial and bifacial minimodules tested for soiling [128], (c) Soiling loss on monofacial and bifacial modules for tilted and vertical mounting using a ratio of short circuit current and global net irradiance [129]. Here, SM, AR, and WB denote different types of front/back sheets, and (d) Our calculations on the normalized yield ratio for bifacial and monofacial tilted modules in a desert climate.
Figure 23. (a) Soiling loss experiments for three configurations: vertical bifacial (VB), latitude tilt bifacial (LB), and latitude tilt monofacial (LM) modules [127], (b) Monofacial and bifacial minimodules tested for soiling [128], (c) Soiling loss on monofacial and bifacial modules for tilted and vertical mounting using a ratio of short circuit current and global net irradiance [129]. Here, SM, AR, and WB denote different types of front/back sheets, and (d) Our calculations on the normalized yield ratio for bifacial and monofacial tilted modules in a desert climate.
Energies 16 05224 g023
Figure 24. Effect of soiling on solar-power generation. (A) PV capacity operational by 2018 along with the medium expectation for 2023. (B) Associated soiling levels. (C) Cleaning costs related to the cleaning methods. (D) Standard energy yield in kWh/kWp for typical geographical spots. (E) An optimal number of yearly cleaning cycles (calculated: bars) and a real range of representative yearly cleaning cycles. Arrows show CSP. (F) The estimated monetary losses due to soiling. Adapted from ref. [40].
Figure 24. Effect of soiling on solar-power generation. (A) PV capacity operational by 2018 along with the medium expectation for 2023. (B) Associated soiling levels. (C) Cleaning costs related to the cleaning methods. (D) Standard energy yield in kWh/kWp for typical geographical spots. (E) An optimal number of yearly cleaning cycles (calculated: bars) and a real range of representative yearly cleaning cycles. Arrows show CSP. (F) The estimated monetary losses due to soiling. Adapted from ref. [40].
Energies 16 05224 g024aEnergies 16 05224 g024b
Figure 25. (a) Cost of electricity generation per selected technologies in the Golf region. * Low = price for 300 MW Sakaka solar PV; and High = a conservative assumption based on project data and expert opinion. ** Low = price for 700 MW in Dubai; and High = price for Morocco’s Noor II. *** Low = price for the Hassyan Clean Coal Power Plant; and High = estimate for coal with CCS. **** Estimated range for nuclear power. (b) Targets of sustainable energy. (c) Already operational renewable energy capacity. (d) Renewable power planned additions by country. Reproduced with permission from ref. [152].
Figure 25. (a) Cost of electricity generation per selected technologies in the Golf region. * Low = price for 300 MW Sakaka solar PV; and High = a conservative assumption based on project data and expert opinion. ** Low = price for 700 MW in Dubai; and High = price for Morocco’s Noor II. *** Low = price for the Hassyan Clean Coal Power Plant; and High = estimate for coal with CCS. **** Estimated range for nuclear power. (b) Targets of sustainable energy. (c) Already operational renewable energy capacity. (d) Renewable power planned additions by country. Reproduced with permission from ref. [152].
Energies 16 05224 g025
Table 1. Experimentally measured adhesion forces between dust particles and glass substrate in OTF (Doha, Qatar) [41].
Table 1. Experimentally measured adhesion forces between dust particles and glass substrate in OTF (Doha, Qatar) [41].
Type of Adhesion ForceMeasured Value (Averaged over Samples)
Capillary1951 nN
Van der Waals39.4 nN in humid conditions, 324 nN in dry air
Electrostatic0.026 nN
Gravitational0.0018 nN
Table 2. Soiling reduction potential and costs for selected soiling mitigation technologies. Adapted from references [84,85].
Table 2. Soiling reduction potential and costs for selected soiling mitigation technologies. Adapted from references [84,85].
Mitigation TechnologyPotential Optimum Reduction of Soiling RatesCostsPotential LimitationsApplication Scenario
Fully automated cleaning>95%2.4–8.2 €/m2Integration in plant design PV utility scale, ground mounted
Anti-soiling coatings:
-
Applied by glass manufacturer
-
Retro-fit
<<80% (literature review)
<20%–50% (authors estimate)
32% reported for
commercial coating
<2 €/m2Performance dependent on location and season, degradation by cleaning and environmental stressesUtility scale, residential, ground mounted and rooftop, BiPV, CSP
+ extra benefit from AR property
Tracking<40%–60%N.A.Integration in plant process; involves additional costsUtility scale, ground mounted, state of the art in CSP
Electrodynamic screen/shield <<98% (laboratory)
32% reported for 2-year study in Saudi Arabia
<30 €/m2Expensive, large-scale application needs to be provenBiPV, island systems, street lighting, rooftop, CSP
Heating
-
PCM
-
Active cell heating
-
PVT
<20%–60%<80 €/m2 (PCM)
N.A.
Expensive, large-scale application needs to be developedBiPV, island systems, street lighting, rooftop installations + extra benefit from cooling during day for PCM + PVT
Optimized PV module design and orientation<65%%0 €/WpIntegration into mass productionUtility scale, rooftop installations
Site adaptionunknown, site specificN.A.Low experience, research neededUtility scale PV and CSP
Table 4. Summary of soiling loss studies for bifacial and monofacial modules.
Table 4. Summary of soiling loss studies for bifacial and monofacial modules.
Study ReferencePV TechnologyTilt AngleSoiling Parameters
Ullah et al. [131]Bifacial30°1.12% (daily soiling loss)
Vertical0.22% (daily soiling loss)
Monofacial30°0.84% (daily soiling loss)
Vertical0.11% (daily soiling loss)
Bhaduri and Kottantharayil [133]Bifacial Vertical0.027%/day
Latitude Tilt0.39%/day
MonofacialLatitude Tilt0.40%/day
Qasem et al. [1]MonofacialVariable Tilt9 times less soiling on vertical modules than 30° tilted modules
Luque, Antonanzas-Torres, and Escobar [134]BifacialLatitude Tilt0.236%/day (Total Soiling rate)
0.0394%/day
(Rear Side Bifacial Soiling rate)
Monofacial0.301%/day (Soiling rate)
Rabanal-Arabach et al. [132]Bifacial VerticalNo notable soiling loss, ~−0.0%
Latitude Tilt−12.5% (power loss due to dust)
MonofacialLatitude Tilt−17.25% (power loss due to dust)
Dassler et al. [135]Bifacial 22° Tilt0.61%/day (Soiling rate)
Monofacial0.57%/day (Soiling rate)
Moehlecke et al. [138]Monofacial and BifacialLatitude TiltPower degradation due to dust is similar for both technologies: 1–4%
Table 5. Predicted renewable energy capacity in GCC countries by 2030. Adapted from ref. [152].
Table 5. Predicted renewable energy capacity in GCC countries by 2030. Adapted from ref. [152].
WindPV Roof-TopPV Utility ScaleCSPWaste to EnergyTotalSource
Capacity in 2030 (MW)
Oman121099024207701105500Target: 2.6 GW (~2025) + 0.6 GW every year up to 2030
Bahrain20705207020700IRENA expectation
Kuwait200100058001000-8000Inputs from country
Qatar-15022506001003100IRENA expectation
UAE300420018,900600060030,000Based on Masdar Institute/IRENA
KSA350075010,500950075025,000Target: 9.5 GW (~2023) + 2 GW every year up to 2030
Table 6. Summary of the main findings.
Table 6. Summary of the main findings.
SectionMain Findings/RecommendationsReferences
Impact of the soiling on solar radiation in Qatar
Development of a national network of 13 high-quality solar radiation-monitoring stations across the state of Qatar.
Measuring the three components of solar radiation (direct, global, or total, and diffuse).
Study of atmospheric dust accumulation on sensor surfaces. [10,11,12,13,14,15,16,17,18,19]
Higher values of the integrated backscatter were seen in the summer season and during periods of dust storms indicating higher aerosol loads.
A clear relationship was pointed out relating to the hourly DNI clearness index.
AOD was well correlated with the particulate matter (PM) mass concentration using spectral ground measurements based on an MFRSR.
Solar radiation and dust accumulation effect on pyranometers in Qatar of the soling ratio quantified (loss per day between 0.3% and 0.5%) depending on the season.
[10,11,12,13,14,15,16,17,18,19]
Development of an in-house “outdoor soiling microscope”
Development of an in-house device “outdoor soiling microscope” consisting of an adapted digital microscope, connected to a computer. A LED and a sheet of translucent paper were used to optimize the contrast of the dust particles.
Fast measurement (within a few seconds) of the droplet formation lifetime, including the onset and the vanishing of the condensation. Deposit dust particles that are above 10 µm2 in size, measuring the Dust Particles removal.
Resolution of 2–3 pixels across, translated into the detection of particles of about 2–3 µm in diameter.
[20]
Field measurement of PV soiling by MarsTM and Dust IQ sensors
Demonstration of field-test data measurements of the Mars™ and Dust IQ soiling sensors installed in different geographical locations in Qatar.
A high correlation between these sensors and a PV cell was demonstrated, especially during the cold season.
The mismatch of the soiling ratio measured from the two methods, namely the Mars™/DustIQ and the conventional clean/soiled reference cell pair method, was found to be less than 0.11%.
[21]
Understanding of the soiling properties in the desert environment
The mean size might be considered at around 20 μm with a dominant spherical-like shape.
The dominant chemical composition of these desert dust particles are (CaSO4), (Ca2Mg(SiO7)), wuestite (FeO), (Al2(SiO4)O), (Mg2(SiO4)), (CaCO3) and (SiO2).
The concentration of airborne dust was found to be a key factor in forecasting the soiling rates over different geographical locations over medium to long periods.
Three parameters were found to constitute the best PV soiling predictors, namely (i) the airborne dust concentration, (ii) the duration of the dry periods, and (iii) the rainfall frequency.
The adhesion forces between the flat glass surface and the dust particles were dominated by capillary forces in the presence of moisture, which may also prevent their resuspension by the wind.
van der Waals forces dominated the adhesion in the case of dry conditions, though gravity and electrostatic forces could be considered negligible.
Under windy conditions, when drag forces were present, rolling was the dominant detachment mechanism for particles.
A threshold of minimum rainfall water to clean the PV module was found to be dependent on the velocity of the droplet, dust composition, surface wettability, tilt angle, and dust adhesion forces, and varied from a daily minimum of 0.3 mm up to 20 mm, depending on the climate factors.
Temperature (T), relative humidity (RH), wind speed (WS), and the concentration of airborne dust, vary along the day as follows: T is lower during the night period, and higher during the daytime. PV module temperature is lower during the night period, and higher during the daytime. RH is higher during the night period, and lower during the day. WS is lower during the night period (around 3 m/s) and higher during the day (around 5 m/s). PM10 concentration has higher mean levels during the night period than during the day.
Surface properties play a critical role in the adhesion of the dust particle.
Typically, the nature of the adhesion forces is surface’ property dependent, and has several forms:
Rough and smooth glass surface: van der Waals forces.
Hydrophobic and hydrophilic surface: capillary forces.
Charged particles: Electrostatic forces.
Gravity, when assuming the weight of the particle sphere.
Under high relative humidity, the adhesion process is governed by capillary force (98%), while during dry conditions, van der Waal force dominates. The main findings are detailed below:
Gravitational forces are negligible for particles size below 500 µm.
PV modules installed in OTF (Qatar) showed 5.1 K degrees below ambient temperature during the night.
The successive drying of the PV panel after the dew period increased dust particles’ adhesion, by cementation, caking, and capillary aging.
Soiling occurs principally during the night period.
In ambient conditions, water droplets were formed on the surfaces of both glass and PTFE for 4 h, at nighttime.
Three processes govern the rate of dust accumulation on a given surface, namely:
“Deposition” refers to the particle in the atmosphere impacting the surface.
“Rebound” is when this particle is rapidly rebounding from this surface without adhering.
“Resuspension” occurs once this particle is resuspended by the wind.
Inclining the PV surface toward the airflow should promote the resuspension of particles.
Heating the PV module to mitigate dew depositions can effectively reduce the cementation process.
Condensation usually occurs before sunrise when the ambient temperature is higher than the module temperature due to radiative cooling, and when the relative humidity is high.
On days of dew prevalence, the soiling levels are recorded to be considerably higher than on dry days.
Active heat with reasonably high energy suggested a soil reduction of up to 65%.
No financial viability study of these warming strategies is yet available.
[22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53]
Impact of PV soiling on PV performance
Sedimentation is affected by gravity and is dependent on the tilt angle.
Inertial and resuspension are affected by wind and are dependent on both tilt angle and azimuth.
Deposition is more important than resuspension, and PV soiling is mainly controlled by sedimentation.
Soiling is more significant on moderately tilted surfaces rather than horizontal ones.
The maximum dust deposition in the OTF field test occurred at 45° facing the wind.
The maximum accumulation occurred at 22° away from the wind.
To decrease the soiling at night, a one-axis PV tracker should be stowed at a maximum tilt angle, facing the wind.
At 5 m/s wind speed, the average median diameter of the accumulated dust particles measured on the coupons was 20 µm. The accumulation of the dust particles was found to decrease with respect to WS and with respect to the tilt angle too.
For PV trackers that rotate commonly around ±45° or ±60°, a tangible reduction of dust accumulation could be achieved “for free,” especially during windy nights, by keeping the PV modules at a maximum tilt (ideally vertically) facing the wind.
Depending on the dominant meteorological conditions, the monthly average daily dust accumulation rate (DAR) was found to vary from 75 to 250 mg/m2/day.
The lowest value of DAR is observed during June and is attributed to the high WS, and low RH% (although higher PM10 was observed during this month).
A DAR of 75 mg/m2/day corresponded to a relative PV power loss of about 0.20%.
The average daily DAR generally follows the same pattern as the airborne PM10 concentration.
Higher DAR favors higher dust accumulation, and hence higher PV soiling.
The summer season (July–October) showed a decrease in the cleanness index (CI) by about 20% per month and this occurs consistently during only dry periods year after year.
Typically, for the bi-monthly cleaned panels, the associated PV power loss was about 15%/ month, with a clearly defined seasonal pattern.
During wet seasons, i.e., rainy periods, the CI reduction was limited to 10% due to cleaning triggered by the rain that prevents the accumulation of dust for more extended periods.
Wet season with higher rain rates such as in 2017 has shown a better CI (even close to unity) as compared, for example, to the same season in 2015.
Although the monthly PV soiling varied noticeably, the associated seasonal trends were more or less respected and the yearly change was less significant.
A threshold of a minimum of 3 mm rainfall was found to be the required value to fully clean the PV modules.
Dust storm days increase the annual average soiling rate by 23%.
[8,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78]
Mitigation of PV soiling
No cleaning process is currently universally recommended, as it depends on the availability of on-site resources, the economics, and the required frequency of cleaning.
Dry-cleaning-based technologies are applied in desert and arid environments where water sources are rare.
Wet-cleaning-based technologies are usually favored owing to their efficiency and the low potential for surface damage.
In the past few years, almost 100% of robotic cleaning was implemented in MW-scale PV projects.
The potential cost for applying automated cleaning is in the range of 2.4–8.2 EUR/m2 with a possible soiling reduction of more than 95%.
The soiling level in the field declined as much as 80% using anti-soiling coatings when applied for a short time period. For extended times, typical anti-soiling efficiency is usually much weaker (for instance 20% –50%) and may grow worse than bare front cover depending on the quality of coating, regional weather, and deterioration level.
The long-term stability and reliability of passive coatings are difficult to foresee.
41% soiling reduction was achieved on average by stowing vertically (90°) at night and a further 9% reduction by stowing upside-down (180°).
[40,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139]
Effect of soiling on bifacial PV modules
Soiling on bifacial PV systems depends on tilt angle, mounting height, and rear/front side dust deposition rates.
Vertical bifacial modules show great potential for free soiling (0.027%/day).
[1,127,128,129,130,131,132,133,134,135,136,137,138]
Impact of soiling on energy cost and power production
The global solar-power production reduction to soiling is estimated to be at least 4% (in 2018), occasioning a conservative estimated monetary loss of at least EUR 5 billion.
By 2023, the worldwide yearly power production losses due to soiling issues may rise to 7%, generating more than EUR 7 billion economic.
[40,136]
Renewable Energy and soiling within the Gulf Cooperation Council (GCC) context
MENA region has the biggest PV generation potential in the world thanks to the abundant sunlight. It is also the stronghold of harsh climate factors, especially the events of dust storms and high particulate matter concentrations.
Soiling considerably decreases the light transmitted to the PV solar cells and hence lowers the output PV power.
The state of Qatar has set a goal of attaining 20% of its energy from solar power by 2030.
One of the largest PV plants in the world, namely Al-Kharsaah, with 800 MWp is already commissioned.
Two more projects accounting for 875 MWp will be delivered by the end of 2024.
Led by the UAE, the GCC region has about 7 GW in renewable power generation capacity.
Solar PV remains the dominant technology with more than 75% share, followed by 10 % of CSP and a 9% share for wind projects.
[139,140,141,142,143,144,145,146,147,148,149,150,151,152]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Aïssa, B.; Isaifan, R.J.; Figgis, B.W.; Abdallah, A.A.; Bachour, D.; Perez-Astudillo, D.; Sanfilippo, A.; Lopez-Garcia, J.; Bermudez Benito, V. A Comprehensive Review of a Decade of Field PV Soiling Assessment in QEERI’s Outdoor Test Facility in Qatar: Learned Lessons and Recommendations. Energies 2023, 16, 5224. https://doi.org/10.3390/en16135224

AMA Style

Aïssa B, Isaifan RJ, Figgis BW, Abdallah AA, Bachour D, Perez-Astudillo D, Sanfilippo A, Lopez-Garcia J, Bermudez Benito V. A Comprehensive Review of a Decade of Field PV Soiling Assessment in QEERI’s Outdoor Test Facility in Qatar: Learned Lessons and Recommendations. Energies. 2023; 16(13):5224. https://doi.org/10.3390/en16135224

Chicago/Turabian Style

Aïssa, Brahim, Rima J. Isaifan, Benjamin W. Figgis, Amir A. Abdallah, Dunia Bachour, Daniel Perez-Astudillo, Antonio Sanfilippo, Juan Lopez-Garcia, and Veronica Bermudez Benito. 2023. "A Comprehensive Review of a Decade of Field PV Soiling Assessment in QEERI’s Outdoor Test Facility in Qatar: Learned Lessons and Recommendations" Energies 16, no. 13: 5224. https://doi.org/10.3390/en16135224

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop