Next Article in Journal
Measurements and Analysis of Polycyclic Aromatic Hydrocarbons near a Major Interstate
Previous Article in Journal
Satellite Assessments of Tropopause Dry Intrusions Correlated to Mid-Latitude Storms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on Predicting Ground PM2.5 Concentration Using Satellite Aerosol Optical Depth

1
Department of Epidemiology and Biostatistics, School of Public Health, Wuhan University, 115# Donghu Road, Wuhan 430071, China
2
Global Health Institute, Wuhan University, 115# Donghu Road, Wuhan 430071, China
3
The National Environmental Satellite, Data, and Information Service (NESDIS), National Oceanic and Atmospheric Administration (NOAA), 5830 University Research Court, College Park, MD 20740, USA
4
International Baccalaureate Diploma Program, Wuhan Foreign Languages School, Wan Song Yuan Road, Wuhan 430022, China
5
Environmental Health Laboratory, Department of Public Health Sciences, University of Hawaii at Manoa, 1960 East-West Road, Honolulu, HI 96822, USA
6
State Key Laboratory of Information Engineering in Surveying, Mapping and Remote Sensing, Wuhan University, 129# Luoyu Road, Wuhan 430079, China
7
College of Information Science and Engineering, Wuchang Shouyi University, Wuhan 430064, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Atmosphere 2016, 7(10), 129; https://doi.org/10.3390/atmos7100129
Submission received: 25 July 2016 / Revised: 26 September 2016 / Accepted: 5 October 2016 / Published: 14 October 2016

Abstract

:
This study reviewed the prediction of fine particulate matter (PM2.5) from satellite aerosol optical depth (AOD) and summarized the advantages and limitations of these predicting models. A total of 116 articles were included from 1436 records retrieved. The number of such studies has been increasing since 2003. Among these studies, four predicting models were widely used: Multiple Linear Regression (MLR) (25 articles), Mixed-Effect Model (MEM) (23 articles), Chemical Transport Model (CTM) (16 articles) and Geographically Weighted Regression (GWR) (10 articles). We found that there is no so-called best model among them and each has both advantages and limitations. Regarding the prediction accuracy, MEM performs the best, while MLR performs worst. CTM predicts PM2.5 better on a global scale, while GWR tends to perform well on a regional level. Moreover, prediction performance can be significantly improved by combining meteorological variables with land use factors of each region, instead of only considering meteorological variables. In addition, MEM has advantages in dealing with the AOD data with missing values. We recommend that with the help of higher resolution AOD data, future works could be focused on developing satellite-based predicting models for the prediction of historical PM2.5 and other air pollutants.

1. Introduction

According to the World Health Organization’s report in 2014, 3.7 million premature deaths related to ambient air pollution occurred around the world in 2012 [1]. Ambient air pollutants include particulate matter, ozone, nitrogen dioxide, sulfur dioxide, and other contaminants. Fine particulate matter with aerodynamic diameters smaller than 2.5 μm (PM2.5) is the most problematic of these pollutants. PM2.5 particles can enter into the alveoli, subsequently being retained in the lung parenchyma [2]. Due to the toxicological effects of the resulting inflammation and oxidative stress [3], PM2.5 can cause severe cardiovascular diseases, respiratory diseases and even lung cancer [4,5]. A study of the global burden of disease study in 1990–2010 ranked ambient PM2.5 concentrations ninth out of all health risk factors [6]. PM2.5 has therefore played an important role in the area of air pollution and environmental health [7,8,9,10].
However, most pollutant concentration information was obtained from ground monitoring stations, which have many limitations. These stations are limited in number, unequally distributed [7,11] and have different measure frequency ranges [12]. These limitations may affect the geographical and demographical range of studies, resulting in an information bias and reducing the confidence in the results of exposure response studies [13]. Furthermore, the temporal and spatial variation of PM2.5 is complex, and continuous monitoring of PM2.5 is absent in many countries and regions [14]. For example, PM2.5 was not included in China’s national monitoring system until 2013. Remote sensing techniques could therefore allow the collection of long period continuous PM2.5 data on large spatial scales over China [15].
Numerous researchers have attempted to estimate ground PM2.5 levels using satellite-derived atmospheric aerosol optical depth (AOD) [16], which is the aerosol extinction coefficient of accumulated points in the vertical direction [4,16,17]. Satellite-derived AOD research began in the mid-1970s, and, in 2003, Wang et al. [16] initiated the use of Moderate Resolution Imaging Spectrometer (MODIS) AOD in the prediction for ground level PM2.5 though linear correlation. Liu et al. [18] came up with Chemical Transport Model (CTM) in 2004, and, in 2011, Lee et al. [19] created the day-specific Mixed-Effect Model (MEM) using MODIS AOD. In recent years, PM2.5 levels have been estimated using a variety of satellite sensors, including the MODIS [20,21], the Multi-Angle Imaging Spectrometer (MISR) [4,20,22], the Geostationary Operational Environment Satellite (GEOS) [23,24], Polarization of Earth’s Reflectance and Directionality (POLDER) [25,26], the Sea-viewing Wide Field-of-view Sensor (SeaWiFS) [27,28], the Ozone Monitoring Instrument (OMI) [29] and the Cloud-Aerosol Lidar with Orthogonal Polarization (CALIOP) [29,30]. Although studies of this kind are becoming more common, prediction results have been unstable and varied significantly between different regions [31,32]. Additionally, different studies have used different methods of dealing with missing AOD data [7,33,34,35,36]. The objective of this study is to review previous studies in order to compare existing PM2.5 predicting models based on satellite AOD and illustrate their advantages and limitations. This could provide a helpful reference for future satellite-based PM2.5 predicting studies.

2. Methods

2.1. Subject of This Review

What is the relationship between PM2.5 concentrations predicted from aerosol optical depth retrieval and PM2.5 concentrations measured on the ground?

2.2. Search Criteria

We searched the following electronic databases prior to 30 June 2016: Web of Science (WOS), PubMed, Engineering Index (EI), Nature, Elsevier Science Direct, Wiley, Springer, and Taylor and Francis. Keywords used in the searches included: aerosol optical depth (AOD, aerosol optical thickness, AOT), fine particulate matter (PM2.5), satellite data, satellite remote sensing, satellite derived, and satellite retrieved. These keywords were searched under the categories of “subject”, “title”, and “keywords” respectively, connected through logical combinations of “and” and “or”. When searching in Web of Science, for example, we used the following combination of keywords: ((“aerosol optical depth”) OR (“AOD”) OR (“aerosol optical thickness”) OR (“AOT”) OR (“satellite data”) OR (“satellite remote sensing”) OR (“satellite derived”) OR (“satellite retrieved”)) AND ((“fine particulate matter”) OR (“PM2.5”)).

2.3. Inclusion and Exclusion Criteria

The inclusion criteria are as follows: (1) papers published in the peer-reviewed journals before 30 June 2016; (2) empirical research utilizing both satellite AOD data and ground PM2.5 data; and (3) papers incorporating PM2.5 predicting models based on satellite-derived AOD and model evaluation. During the process of abstract and full texts reviewing, studies were excluded according to these criteria: (1) abstracts and conferences only; (2) studies using AOD data only or PM2.5 data only, and studies without R2 values; and (3) satellite-based PM2.5 predicting studies conducted over the ocean or special terrains (such as mountains), or during the following natural and anthropogenic events: land (forest) fires, dust storms, volcanic eruptions, and fuel combustion events. We reviewed all the selected studies in detail and summarized their main features.

3. Results

After screening 1436 identified studies and assessing the eligibility of the remaining studies, we selected 116 articles for our review that are primarily relevant to the satellite-based PM2.5 predicting model (Figure 1). The study areas, results, models used and other basic characteristics of all included studies are summarized in Table 1.
Of these 116 studies, 25 used Multiple Linear Regression (MLR), 23 used the Mixed-Effect Model (MEM), 16 used the Chemical Transport Model (CTM), and 10 used Geographically Weighted Regression (GWR), while Linear Correlations (LC), the Generalized Additive Model (GAM), Land Use Regression (LUR) and others models were found in 12 studies, six studies, seven studies and 27 studies, respectively (Figure 2 and Figure 3).

4. Discussion

Satellite remote sensing technology plays an essential role in the field of meteorology because of its highly accurate prediction of meteorological disasters. Recently, this technology has also been used in the prediction of daily air pollution (PM2.5) levels. Although PM2.5 data can be obtained from AOD measured by ground-based remote sensing equipment [129], it is more meaningful to predict PM2.5 levels from satellite observations. From our study, we concluded that MLR, MEM, CTM, and GWR were the models most commonly used to predict PM2.5 levels.

4.1. Multiple Linear Regression

4.1.1. Theory Background and Application

MLR has been used to predict PM2.5 from satellite AOD since 2005. In this model, PM2.5 measured at ground level PM2.5 was set as the dependent variable, and AOD was set as the independent variable. Several factors were also included in the model as covariates, including humidity, temperature, wind speed, wind direction, aerosol type, and height of the boundary layer. MLR was often used in earlier studies to predict PM2.5 levels. For instance, Liu et al. [38] used this model to analyze three area types (city, and suburb and countryside) in the eastern United States in 2001. They reported that coefficients were quite low to some extent and also varied greatly between different regions; R2 values were 0.420, 0.490, 0.590 and 0.430 in city, suburban, countryside and whole area, respectively. The low R2 value showed above indicated that the inclusion of covariates (such as relative humidity, height of the boundary layer, season variable, etc.) in MLR models requires further discussion [33]. In contrast, R2 value reached up to 0.960 in Gupta’s study when certain conditions (weather condition, boundary layer heights and others conditions) were met [41].
More recently, in order to improve model performance, some studies have explored covariate factors in the MLR model under different conditions [17,20,21,24,25,26,39,40,41,46,47,49,50,54,75]. A few covariate factors, such as relative humidity and height of the boundary layer, were regarded as significant enough to affect and even invert the relationships between AOD and PM2.5. In 2013, Cordero et al. [73] predicted PM2.5 levels by applying both the satellite-based MLR method and the Community Multi-Scale Air Quality (CMAQ) model. Results showed that the satellite-based MLR method performed better than the CMAQ model during summer: R2 values ≥0.423 (MODIS), R2 values ≥0.137 (CMAQ). However, the R2 value increased to 0.740 when the two models were combined [73]. In 2015, Han et al. carried out affecting factors analyses between AOD and PM2.5 in Nanjing [93]. The authors found that aerosol type and height of the boundary layer were significant factors in the prediction of PM2.5 levels. They also stated that R2 value was 0.624 with only aerosol type adjusted, and R2 value was 0.548 when both aforementioned significant factors were adjusted [93].

4.1.2. Advantages and Disadvantages

In summary, the determination coefficients of MLR were relatively higher than those of the linear correlation model, and a confounding bias could be avoided by including relevant covariates into the model. However, there are several limitations. Some important covariates, such as seasonal variation of the aerosol, regional variation, and land use information, were missing from the models [93]. Additionally, the accuracy and resolution of the satellite-derived AOD and meteorological data was low [38], which can lead to an information bias.

4.2. Mixed-Effect Model

4.2.1. Theory Background and Application

In early research, missing AOD data was an essential factor in the estimation of PM2.5 from AOD, and the method used to compensate for missing AOD data is a very important factor in the precision and accuracy of the derivation. Kloog et al. [33], from Harvard School of Public Health, first proposed that satellite-derived AOD could be included in the three-stage MEM and they applied this approach in New England in 2011. Based on the AOD day-specific correction mixed-effect model of Lee et al. [19], they took meteorological variables and classic land use variables into the MEM [34]. The MEM also used the inverse distance weight method (IDW), cluster analysis, GAM and generalized additive mixed model (GAMM) to deal with missing AOD values so that daily ground PM2.5 levels could be predicted in a wide range [34]. If missing AOD presented non-random distribution, AOD data needed to be corrected by meteorological factors using the inverse probability weight method (IPW) [82].
MEM has been applied in many regions. In New England, Kloog et al. [33] constructed their own MEM, based on MEM of Lee et al., in 2011 (CV R2 value = 0.830, for days with available AOD data; 0.810, for days without AOD data). The distinctive feature of the model of Kloog et al. is its inclusion of meteorological variables (such as temperature, wind speed and visibility) and land use variables (such as elevation, percentage of open spaces, area emissions, point emissions and distance to major roads) into the model, which is appropriate for studying acute and chronic health effects. Since then, many researchers, including Kloog, Madrigano, Chiu and others, have used MEM to study acute and chronic health effects [79,96,130,131,132,133], and it has performed well. In 2012, by using GEOS AOD data and adding a surface reflection variable into MEM, Chudnovsky et al. [35] showed a high predictive value of CV-R2 = 0.920. This study also proved that high resolution GEOS AOD may be a better predictor of urban PM2.5 than rough resolution MODIS AOD [35]. Lee et al. [64] found that the R2 value of MEM could reach 0.830 if missing AOD value were filled using a combination of cluster analysis and generalized additive models. In the mid-Atlantic region, Kloog et al. [34] improved the MEM by adopting IPW for non-random missing AOD data, and obtained a value of 0.850 for the cross-validation of R2. Kloog et al. [34] also established PM2.5 predictive models in different regions by adding traffic density, population density and distance to the point emission variables.
In 2013, in order to take advantage of high resolution AOD products. Chudnovsky et al. [56] developed a MODIS based Multi-Angle Implementation of Atmospheric Correction algorithm (MAIAC) and used this new algorithm to improve the inversion resolution of MODIS AOD products (from 10 × 10 km to 1 × 1 km). In their results, R2 value reached up to 0.500 in New England and 0.860 in Boston area [72]. The MAIAC algorithm has since been widely applied in MEM studies [7,82,83,94,120,134]. Kloog et al. [83] obtained a CV-R2 value of 0.810 in mid-western United States in 2000–2006. In a later study based on an early MEM [33,34], Kloog et al. performed a GAM to address missing AOD values, obtaining an R2 value of 0.880 in the northeastern United States (New England, New York and New Jersey) [82]. In New England, Alexeeff et al. [135] further employed the MEM model with [34,131] Kriging and land use regression to describe an epidemiological relationship between AOD and predicted PM2.5 in 2003. The following year, Shi et al. [120] used MEM to predict PM2.5 using MODIS AOD data collected between 2003 and 2008, and they obtained consistent results (R2 value = 0.890) for days with available AOD data and without available AOD data. This method was also successfully applied in studies on the relationship between low PM2.5 exposure and mortality.
In recognition of regional geographical differences, Lee et al. [7] predicted PM2.5 concentrations using IPW in seven southeastern states of the United States in 2016, and they obtained three coefficients of determination (0.770, 0.810, and 0.700) from three different geographical area types. They suggested that their PM2.5 estimation methods could be applied from urban areas to rural areas. Just et al. [94] analyzed the geographical distribution of PM2.5 in Mexico in 2004–2014. They obtained an R2 value of 0.724 using MEM and showed that precipitation and height of the boundary layer are both important factors influencing the relationship between AOD and PM2.5 [94]. Furthermore, with AOD derived from Medium Resolution Imaging Spectrometer (MERIS) and Advanced Along-Track Scanning Radiometer (AATSR) synergistic observations. Beloconi et al. [108] applied MEM to the evaluation of the day-specific and site-specific random effects in London. Their results showed a CV-R2 value of 0.846 between 2002 and 2012. Ma et al. [136] provided an improved MEM to address data missing from satellite observation as well as ground-level measurements.

4.2.2. Advantages and Disadvantages

To sum up, MEM had the following advantages: (1) It had a relatively high predicting coefficient of determination. The R2 value could generally reach up to 0.800 or higher. R2 values of time and spatial consistency were also high; they could reach up to 0.700 or higher among different regions. Besides, R2 value could be greatly enhanced through the use of MAIAC algorithms [7,72,82]; (2) MEM could be widely applied to the prediction of PM2.5 at a regional level by using different land use and meteorological variables for model calibration; (3) MEM can be used to predict daily PM2.5 concentrations, and has been applied in studies on the acute and chronic health effects of PM2.5 exposure in New England, the Mid-Atlantic and other regions of the United States. These studies can be extended to other regions in the future [15]. The model may also be used to explore the difference between satellite-derived AOD-based PM2.5 data and ground based PM2.5 data in health effect studies.
MEM has the following disadvantages: (1) Due to the lack of ground-level PM2.5 monitoring data in certain areas, the PM2.5 monitoring data could not meet the requirements of Kriging in MEM, which affected the accuracy of the results [72]; (2) The determination of correlation between AOD and PM2.5 may decrease when only total AOD is applied. It is not clear which of the aerosols influencing AOD (such as sulfate, nitrate, ammonium, carbonaceous, mineral dust, and sea salt) plays a major role in the total AOD, or how much other air pollutants affect this correlation [33,94,137]; (3) Land use and traffic pollution information is hard to collect.

4.3. Chemistry Transport Model

4.3.1. Theory Background and Application

Based on the characteristics of vertical distribution and transmission of AOD, Liu et al. [18] proposed the Global atmospheric chemistry model (GEOS-CHEM), which is a prediction model of PM2.5 based on satellite AOD. Following Liu and coworkers’ study [138], van Donkelaar et al. [43] developed the CTM which calibrates the height of the boundary layer and the humidity of air. Considering the composition and distribution of AOD and utilizing emissions listing data as well as daily emission patterns published in European and other countries, van Donkelaaar et al. built a precise CTM formula in 2006. In 2010, they simplified the CTM by redefining the association between AOD and PM2.5 as a conversion factor. CTM can now be used on a global as well as a local scales [6,57], and has attracted extensive interest [139,140].
This model was employed in different regions between 2010 and 2012. Di Nicolantonio and Cacciari [55] applied the method in North Italy and obtained different results for satellite-based PM2.5 predicting results (R2 values of 0.680 (Terra MODIS), 0.590 (Aqua MODIS), 0.700 (Terra and Aqua MODIS, respectively). Hystad et al. [62] obtained an R2 value of 0.410 for the first time to add land use variables in Canada using CTM [57]. Additionally, in a comparison between IDW-adjusted CTM and MLR, IDW-adjusted CTM (R2 value = 0.510 per year) performed better than MLR (R2 value = 0.330 per year) [62,67]. Lee et al. [63] made a comparison between the Kriging method and CTM in the United States. Although both methods gave consistent results, CTM had better applicability and higher accuracy, especially in areas with few ground level monitoring sites. Further studies by van Donkelaar et al. have shown that meteorological factors can calibrate and reduce the system error and spatial smoothing of the IDW method can reduce the random error, eventually extending the spatiotemporal prediction scale [67]. Crouse [141] not only obtained a high R2 value (0.792) in 11 Canadian cities in 1987–2001, but also successfully applied their results to the study of long-term health effects of PM2.5 exposure. Following van Donkelaar’s study [57], others studies conducted by Villeneuve, Chen, To and Brauer [142,143,144,145,146] focused on acute and chronic health effects and on the global burden of disease.
In addition, the estimates of PM2.5 from MODIS AOD in the above studies were somewhat varied. In 2013, van Donkelaar et al. [147] added land use type data, which were used to quantify the weight of AOD data, and proposed Optimal Estimation (OE) in order to improve the predictive ability of AOD. More recently, Wang et al. [124] have provided an improved AOD retrieval algorithm for MODIS at 1 km resolution that can be retrieve AOD at high spatial resolutions at intra-urban scales. These MODIS-retrieved AODs are used to predict ground level PM2.5 using aerosol vertical profiles and local scale factors obtained from the CTM simulation. Daily R2 value = 0.860 and monthly R2 value = 0.930 were obtained from data collected over the city of Montreal, Canada [124].
At the global level, in a study similar to van Donkelaar’s 2010 study, Boys and Martin [148] completed a global ground level prediction of PM2.5 in 2014, which integrated global AOD data collected from the MISR and SeaWiFS AOD (1 km × 1 km) satellite sensor between 1998 and 2012. They also included a few effecting factors in the CTM, such as the vertical structure of aerosol extinction, relative humidity, aerosol size and component of aerosol variables. Their results showed that PM2.5 levels in East America, the Arabian Peninsula, Eastern and southern Asia were relatively consistent [148]. In a different study, van Donkelaar et al. [101] combined GWR with CTM, and obtained a higher value of CV-R2 (0.780) with high resolution in North America. In the same year, van Donkelaar et al. [27] improved the CTM approach to the prediction of PM2.5 concentrations at a global level. Their research integrated AOD data from three satellites in order to avoid negative effects from the source variations of AOD. The study obtained high R2 values (0.656) for North America in 2001–2010, indicating that PM2.5 prediction could be feasible at the global level.

4.3.2. Advantages and Disadvantages

Based on above studies, the advantages of CTM are: (1) it can predict PM2.5 concentrations at ground level without PM2.5 data from ground monitors [127]; and (2) it takes the component of AOD and the effects of other pollutants into account, and has been widely used in Canada, North America and South America, for predicting on a global scale [27,28,149,150]. CTM is currently central to health effect analysis related to PM2.5 components [109]. The disadvantages of CTM are: (1) the prediction effect was relatively low and variant among different regions. Considering the poor performance of CTM, lower R2 values can lead to a high exposure bias in health effect studies; (2) it will consume time, energy and financial resources to collect the necessary chemical and physical information on PM2.5 [57]; (3) due to the lack of pollutants emissions type and emissions listing data in developed countries, it is hard to meet the conditions of application of CTM in China, India and other developing countries [27]; and (4) other pollutants (SO2, O3, etc.) have different inversion resolutions compared with PM2.5 [143].

4.4. Geographical Weighted Regression

4.4.1. Theory Background and Application

Based on the assumption that “regression coefficient is a function of the observation point’s spatial position in linear regression” with spatial weight assigned according to the distance between observation points, Geographical Weighted Regression (GWR) was first proposed [151,152]. This spatial regression technique reflects spatial variability and non-smooth character, and could provide a regional-level regression model [151,152,153]. In 2009, Hu et al. [32] introduced AOD into GWR and carried out a prediction of PM2.5 levels in the United States. After that, Ma [87] further optimized GWR in 2014 by taking AOD, land use variables as the independent variables, and PM2.5 concentrations as the dependent variable. Meanwhile, based on the differences between regions in PM2.5 ground monitoring, spatial weight assignment was developed and applied to each region with the quantity of AOD data. If a large proportion of AOD data was missing, we could select certain buffer areas for each spatial observation point and fill in the vacancy according to the corrected Akaike Information Criterion. Thus, spatial distribution of regression parameter gained, and the GWR model could explain the effects of the spatial autocorrelation within a certain area when spatial aggregation occurred for a certain variable [87,105,107,128].
Hu’s initial investigation on GWR found that it had a low R2 value compared with MEM and CTM, probably because not all studies took meteorological factors and land use factors into account [32]. Based on regional differences, Hu et al. [59] brought meteorological variables and land use variables into the GWR to predict PM2.5 concentrations in North America. Results showed that R2 values improved significantly when these variables were considered (R2 = 0.672 (North American Regional Reanalysis data), and R2 = 0.706 (North American Land Data Assimilation System data)). However, large spatial variability and instability occurred in these variables. Further studies showed that PM2.5 concentrations were higher in urban areas, and lower in rural villages or mountain areas.
In order to compensate the basis without considering the cross-validation, Ma et al. [87] expanded the National GWR model with data from the newly built national monitoring network to predict PM2.5 levels in China, reporting a CV-R2 value of 0.640. This result indicated that it was feasible to estimate PM2.5 levels in China using satellite AOD combined with meteorological and land use data. The model obtained similar results to those obtained by the CTM used by van Donkelaar in 2010, but GWR found higher PM2.5 concentrations in rural areas. Similar results for national PM2.5 levels were found by You et al. [126] with CV-R2 values of 0.760 and 0.810 for MODIS and MISR, respectively, in China. Additionally, using 3-km resolution MODIS AOD in 2014, You et al. [125] confirmed that this GWR approach is useful for estimating large-scale ground-level PM2.5 distributions in China.

4.4.2. Advantages and Disadvantages

From the studies above, the advantages of this model are: (1) PM2.5 estimation requires only small amounts of data. For example, this model can work with the daily average, monthly average or yearly average of both PM2.5 data alone or AOD data alone. Determination coefficients were also less affected. Studies have shown that compared with CTM, GWR had a higher R2 value [87]; (2) Similar to MEM, GWR used ground monitored PM2.5 values for AOD calibration, and it had a better model performance than MLR. The disadvantages are: (1) since model construction depends on ground monitoring data, model performance may be much less reliable in areas lacking ground monitoring data; and (2) to our knowledge, GWR has only been employed in limited PM2.5 prediction studies with the combination of satellite data [74,87,100,107,125,126,128], so the feasibility of applying it widely in other regions needs to be investigated in further research.

4.5 Other Models

In addition to the models mentioned above, other researchers used linear correlations [16,30,31,37,42,58,71,113,115,117], GAM [23,24,53,65,77], LUR [66,69,70,78,91,122], Kriging [88,90,108] or the nonlinear regression model. Those PM2.5 estimating models all regard AOD as the primary independent variable. As a result, the predictability of these models was limited. Their R2 values were generally lower, and varied between different areas. However, these listed models have been gradually optimized or integrated into other models, as with artificial neural networks (ANN, which incorporate LUR in the CTM) [52,61,68,110,111] and the two stage model (TSM, which combine the GWR with MEM) [80,81,119,121]. In recent years, with the development of the AOD-based mathematical model, many new methods have been developed, such as geographically and temporally weighted regression (GTWR) [107], support vector regression methods (SVR) [99] and machine learning regression (which is a combination of SVR, Gauss neural network processes, Decision trees, and Random forests) [28]. Although these new methods had been proposed, their reliability and veracity need to be investigated in further studies.

4.6. Summary

In terms of the accuracy of PM2.5 prediction, though no single model can replace all others, some existing models have their advantages in the following areas. (1) Model predictability: MLR was commonly used in early studies [17,20,21,24,25,26,39,40,41,46,47,49,50,54,75], whereas MEM and CTM gradually became the dominant methods and replaced MLR after 2010. However, GWR has developed at a slower pace with a limited number of studies to data, and had moderate performance [32,74,125,126]. Included studies showed that R2 value of MEM was higher than those of the other three models in the same area [17,87,104,136]. Moreover, MAIAC algorithms, which led to a highly accurate of AOD, were mostly used in MEM, significantly improving the R2 value of the model [7,35,83,120,135]. On the global scale, CTM has been proven to be efficient for the mechanism of completing the prediction from using partial AOD data by AOD component analysis [57]; (2) Adjusting factors: The number of these factors has increased due to the development of prediction models. Moreover, factors such as atmospheric boundary layer height and relative humidity have become a permanent part of the adjustment process. In early LC and MLR studies, adjusting factors were limited in number and scope, and were mainly focused on meteorological factors (atmospheric boundary layer height, humidity, temperature, wind speed, etc.) [38,39,41,42]. Later on, GAM took both meteorological factors and land use factors into account, which increased the performances [23,77]. MEM and CTM also incorporated more meteorological factors and land use factors; their R2 values proved to be satisfactory; (3) Missing AOD: Although predicting of PM2.5 with satellite AOD has become the hotspot in remote sensing field, missing values of AOD cannot be ignored, because the predicted reliability of PM2.5 could be affected when the percentage of missing AOD values reach 60%. Among the four models, MEM systematically and comprehensively described methods of dealing with missing AOD [137]; results of each method could be found in different studies. CTM, on the other hand, filled in the vacancy by establishing “buffer areas” or avoided the problem of missing AOD by assigning different weights to each area according to the amount of AOD data. For the MLR, missing AOD was not processed.

5. Conclusions

The review showed that MEM performed best. CTM had strengths in the prediction of PM2.5 on a global scale. GWR was suitable for PM2.5 prediction on a regional scale. MLR was relatively weak in terms of predictability. When land use information was included as an adjustment factor in addition to meteorological factors, the accuracy of predictions greatly improved. Other models, such as ANN, TSM and SVR, need to be further validated. We therefore suggest that the following possibilities be considered in future studies: (1) the use of AOD data with higher resolution for more accurate estimation of PM2.5 in relatively small areas; (2) the use of satellite-based predicting models for historical PM2.5 prediction and retrospective study in areas lacking historical PM2.5 data; and (3) the development prediction models not only for PM but also for other air pollutants (SO2, NO2), to extend the applicability of predicting models.

Acknowledgments

This review was funded by National Natural Science Foundation of China (No. 41571344); the National Key Research and Development Program of China (No. 2016YFC0200900); the Hubei Province Health and Family Planning Scientific Research Project (Grant No. WJ2015Q023); the Fundamental Research Funds for the Central Universities (Grant No. 2042016kf0165); China Postdoctoral Science Foundation (No. 2015M572198); the program of Key Laboratory for National Geographic Census and Monitoring, National Administration of Surveying, Mapping and Geo-information (No. 2014NGCM); Planning Project of Innovation and Entrepreneurship Training of National Undergraduate (No. 201510486102); and Innovation Experiment Program of Medical Students of Wuhan University (No. MS2015037). We express our great thanks to scholars from School of public health, Wuhan University and State Key Laboratory of Information Engineering, Mapping and Remote Sensing, Wuhan University for their helpful suggestion and discussion.

Author Contributions

The study was carried out in collaboration between all authors. Hao Xiang brought idea formation. Yuanyuan Chu, Yisi Liu and Xiangyu Li drafted the manuscript. Zhiyong Liu and Hanson contributed to make the outline, check results of all models and write the discussion part. Xi Chen, Na Li, Meng Ren and Feifei Liu selected articles under the inclusion and exclusion criteria. Yuan’an Lu, Zongfu Mao, Liqiao Tian and Zhongmin Zhu introduced the technologies of deriving PM2.5 from Aerosol Optical Depth.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
AODAerosol Optical Depth
MODISModerate Resolution Imaging Spectrometer
MISRMulti-Angle Imaging Spectrometer
GEOSGeostationary Operational Environment Satellite
SeaWiFSSea-viewing Wide Field-of-view Sensor
POLDERPolarization of Earth’s Reflectance and Directionality
CALIOPCloud-Aerosol Lidar with Orthogonal Polarization
GOCIGeostationary Ocean Color Imager
OMIOzone Monitoring Instrument (OMI)
AATSRAdvanced Along-Track Scanning Radiometer
MERISMedium Resolution Imaging Spectrometer
LCLinear Correlations
MLRMultiple Linear Regression
LURLand Use Regression
GAMGeneralized Additive Model
MEMMixed-Effect Model
CTMChemical Transport Model
GLMGeneral Linear regression Model
GWRGeographically weighted regression
TWRTemporally Weighted Regression
GTWRGeographically and Temporally Weighted Regression
ANNArtificial Neural Networks
SVRSupport Vector Regression
MCAMaximum Covariance Analysis
CMCACombined Maximum Covariance Analysis
TVMTwo-variate method
MVMMultivariate method
OLSOrdinary Least Squares model
TSMTwo-Stage Model
MAIACMulti-Angle Implementation of Atmospheric Correction algorithm
DSADeletion/Substitution/Addition
BMEMBayesian Maximum Entropy method
Nested MEMNested Mixed-Effect Model
Non-nested MEMNon-nested Mixed-Effect Model
SECSurface Extinction Coefficient
BTHBeijing-Tianjin-Hebei region
PRDPearl River Delta region
YRDYangtze River Delta region
NARRNorth American Regional Reanalysis
NLDASNorth American Land Data Assimilation System
Sample-based CV-R2Sample-based Cross Validated-coefficient of determination
DOY-based CV-R2Day-of-Year-based Cross Validated-coefficient of determination

References

  1. World Health Organization. 7 Milion Premature Death in Annually Linked to Air Pollution. Available online: http://www.who.int/mediacentre/news/releases/2014/air-pollution/en/ (accessed on 25 March 2014).
  2. Dockery, D.W. Heath effects of particulate air pollution. Ann. Epidemiol. 2009, 19, 257–263. [Google Scholar] [CrossRef] [PubMed]
  3. Risom, L.; Moller, P.; Loft, S. Oxidative stress-induced DNA damage by particulate air pollution. Mutat. Res. 2005, 592, 119–137. [Google Scholar] [CrossRef] [PubMed]
  4. Hoek, G.; Krishnan, R.M.; Beelen, R.; Peters, A.; Ostro, B.; Brunekreef, B.; Kaufman, J.D. Long-term air pollution exposure and cardio-respiratory mortality: A review. Environ. Health 2013, 12, 43. [Google Scholar] [CrossRef] [PubMed]
  5. Brook, R.D.; Rajagopalan, S.; Pope, C.A., 3rd; Brook, J.R.; Bhatnagar, A.; Diez-Roux, A.V.; Holguin, F.; Hong, Y.; Luepker, R.V.; Mittleman, M.A.; et al. Particulate matter air pollution and cardiovascular disease: An update to the scientific statement from the american heart association. Circulation 2010, 121, 2331–2378. [Google Scholar] [CrossRef] [PubMed]
  6. Lim, S.S.; Vos, T.; Flaxman, A.D.; Danaei, G.; Shibuya, K.; Adair-Rohani, H.; Amann, M.; Anderson, H.R.; Andrews, K.G.; Aryee, M.; et al. A comparative risk assessment of burden of disease and injury attributable to 67 risk factors and risk factor clusters in 21 regions, 1990–2010: A systematic analysis for the global burden of disease study 2010. Lancet 2012, 380, 2224–2260. [Google Scholar] [CrossRef]
  7. Lee, M.; Kloog, I.; Chudnovsky, A.; Lyapustin, A.; Wang, Y.; Melly, S.; Coull, B.; Koutrakis, P.; Schwartz, J. Spatiotemporal prediction of fine particulate matter using high-resolution satellite images in the southeastern U.S. 2003–2011. J. Expo. Sci. Environ. Epidemiol. 2016, 26, 377–384. [Google Scholar] [CrossRef] [PubMed]
  8. Sinha, P.R.; Gupta, P.; Kaskaoutis, D.G.; Sahu, L.K.; Nagendra, N.; Manchanda, R.K.; Kumar, Y.B.; Sreenivasan, S. Estimation of particulate matter from satellite- and ground-based observations over Hyderabad, India. Int. J. Remote Sens. 2015, 36, 6192–6213. [Google Scholar] [CrossRef]
  9. Mordukhovich, I.; Coull, B.; Kloog, I.; Koutrakis, P.; Vokonas, P.; Schwartz, J. Exposure to sub-chronic and long-term particulate air pollution and heart rate variability in an elderly cohort: The normative aging study. Environ. Health 2015, 14, 1–10. [Google Scholar] [CrossRef] [PubMed]
  10. Liu, D.-J.; Li, L. Application study of comprehensive forecasting model based on entropy weighting method on trend of PM2.5 concentration in Guangzhou, China. Int. J. Environ. Res. Public Health 2015, 12, 7085–7099. [Google Scholar] [CrossRef] [PubMed]
  11. Gupta, P.; Christopher, S.A. An evaluation of Terra-MODIS sampling for monthly and annual particulate matter air quality assessment over the southeastern United States. Atmos. Environ. 2008, 42, 6465–6471. [Google Scholar] [CrossRef]
  12. Lee, M.; Koutrakis, P.; Coull, B.; Kloog, I.; Schwartz, J. Acute effect of fine particulate matter on mortality in three southeastern states from 2007–2011. J. Expo. Sci. Environ. Epidemiol. 2015. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, Y.; Koutrakis, P.; Kahn, R. Estimating fine particulate matter component concentrations and size distributions using satellite-retrieved fractional aerosol optical depth: Part 2—A case study. J. Air Waste Manag. Assoc. 2007, 57, 1360–1369. [Google Scholar] [PubMed]
  14. Zhang, Y.L.; Cao, F. Fine particulate matter (PM2.5) in china at a city level. Sci. Rep. 2015, 5, 14884. [Google Scholar] [CrossRef] [PubMed]
  15. Liu, Y. New directions: Satellite driven PM2.5 exposure models to support targeted particle pollution health effects research. Atmos. Environ. 2013, 42, 6465–6471. [Google Scholar] [CrossRef]
  16. Wang, J. Intercomparison between satellite-derived aerosol optical thickness and PM2.5 mass: Implications for air quality studies. Geophys. Res. Lett. 2003, 30. [Google Scholar] [CrossRef]
  17. Tao, J.; Zhang, M.; Chen, L.; Wang, Z.; Su, L.; Ge, C.; Han, X.; Zou, M. A method to estimate concentrations of surface-level particulate matter using satellite-based aerosol optical thickness. Sci. China Earth Sci. 2013, 56, 1422–1433. [Google Scholar] [CrossRef]
  18. Liu, Y. Mapping annual mean ground-level PM2.5 concentrations using multiangle imaging spectroradiometer aerosol optical thickness over the contiguous United States. J. Geophys. Res. 2004, 109. [Google Scholar] [CrossRef]
  19. Lee, H.J.; Liu, Y.; Coull, B.A.; Schwartz, J.; Koutrakis, P. A novel calibration approach of MODIS AOD data to predict PM2.5 concentrations. Atmos. Chem. Phys. Discuss. 2011, 11, 9769–9795. [Google Scholar] [CrossRef]
  20. Liu, Y.; Franklin, M.; Kahn, R.; Koutrakis, P. Using aerosol optical thickness to predict ground-level PM2.5 concentrations in the St. Louis area: A comparison between MISR and MODIS. Remote Sens. Environ. 2007, 107, 33–44. [Google Scholar] [CrossRef]
  21. Zhang, H.; Hoff, R.M.; Engel-Cox, J.A. The relation between moderate resolution imaging spectroradiometer (MODIS) aerosol optical depth and PM2.5 over the United States: A geographical comparison by U.S. Environmental Protection Agency regions. J. Air Waste Manag. Assoc. 2009, 59, 1358–1369. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, Y.; Koutrakis, P.; Kahn, R. Estimating fine particulate matter component concentrations and size distributions using satellite-retrieved fractional aerosol optical depth: Part 1—Method development. J. Air Waste Manag. Assoc. 2007, 57, 1351–1359. [Google Scholar] [PubMed]
  23. Liu, Y.; Paciorek, C.J.; Koutrakis, P. Estimating regional spatial and temporal variability of PM2.5 concentrations using satellite data, meteorology, and land use information. Environ. Health Perspect. 2009, 117, 886–892. [Google Scholar] [CrossRef] [PubMed]
  24. Paciorek, C.J.; Liu, Y.; Moreno-Macias, H.; Kondragunta, S. Spatiotemporal associations between goes aerosol optical depth retrievals and ground-level PM2.5. Environ. Sci. Technol. 2008, 42, 5800–5806. [Google Scholar] [CrossRef] [PubMed]
  25. Leon, J.-F.; Liousse, C.; Galy-Lacaux, C.; Doumbia, T.; Cachier, H. Monitoring of ambient fine particulate matter concentrations from space: application to European and African cities. Proc. SPIE 2010, 78262A. [Google Scholar] [CrossRef]
  26. Kacenelenbogen, M.; Leon, J.F.; Chiapello, I.; Tanre, D. Characterization of aerosol pollution events in France using ground-based and polder–2 satellite data. Atmos. Chem. Phys. 2006, 6, 4843–4849. [Google Scholar] [CrossRef]
  27. Van Donkelaar, A.; Martin, R.V.; Brauer, M.; Boys, B.L. Use of satellite observations for long-term exposure assessment of global concentrations of fine particulate matter. Environ. Health Perspect. 2015, 123, 135–143. [Google Scholar] [CrossRef] [PubMed]
  28. Lary, D.J.; Faruque, F.S.; Malakar, N.; Moore, A.; Roscoe, B.; Adams, Z.L.; Eggelston, Y. Estimating the global abundance of ground level presence of particulate matter (PM2.5). Geosp. Health 2014, 7, S611–S630. [Google Scholar] [CrossRef] [PubMed]
  29. Li, J.; Carlson, B.E.; Lacis, A.A. How well do satellite AOD observations represent the spatial and temporal variability of PM2.5 concentration for the United States? Atmos. Environ. 2015, 102, 260–273. [Google Scholar] [CrossRef]
  30. Toth, T.D.; Zhang, J.; Campbell, J.R.; Hyer, E.J.; Reid, J.S.; Westphal, D.L. Impact of data quality and surface-to-column representativeness on the PM2.5/satellite AOD relationship for the contiguous United States. Atmos. Chem. Phys. Discuss. 2014, 14, 6049–6062. [Google Scholar] [CrossRef] [Green Version]
  31. Engel-Cox, J.A.; Holloman, C.H.; Coutant, B.W.; Hoff, R.M. Qualitative and quantitative evaluation of MODIS satellite sensor data for regional and urban scale air quality. Atmos. Environ. 2004, 38, 2495–2509. [Google Scholar] [CrossRef]
  32. Hu, Z. Spatial analysis of MODIS aerosol optical depth, PM2.5, and chronic coronary heart disease. Int. J. Health Geogr. 2009, 8, 27. [Google Scholar] [CrossRef] [PubMed]
  33. Kloog, I.; Koutrakis, P.; Coull, B.A.; Lee, H.J.; Schwartz, J. Assessing temporally and spatially resolved PM2.5 exposures for epidemiological studies using satellite aerosol optical depth measurements. Atmos. Environ. 2011, 45, 6267–6275. [Google Scholar] [CrossRef]
  34. Kloog, I.; Nordio, F.; Coull, B.A.; Schwartz, J. Incorporating local land use regression and satellite aerosol optical depth in a hybrid model of spatiotemporal PM2.5 exposures in the mid-atlantic states. Environ. Sci. Technol. 2012, 46, 11913–11921. [Google Scholar] [CrossRef] [PubMed]
  35. Chudnovsky, A.A.; Lee, H.J.; Kostinski, A.; Kotlov, T.; Koutrakis, P. Prediction of daily fine particulate matter concentrations using aerosol optical depth retrievals from the Geostationary Operational Environmental Satellite (GOES). J. Air Waste Manag. Assoc. 2012, 62, 1022–1031. [Google Scholar] [CrossRef] [PubMed]
  36. Higgs, G.; Sterling, D.A.; Aryal, S.; Vemulapalli, A.; Priftis, K.N.; Sifakis, N.I. Aerosol optical depth as a measure of particulate exposure using imputed censored data, and relationship with childhood asthma hospital admissions for 2004 in Athens, Greece. Environ. Health Insights 2015, 9, 27–33. [Google Scholar] [CrossRef] [PubMed]
  37. Hutchison, K.D.; Smith, S.; Faruqui, S.J. Correlating MODIS aerosol optical thickness data with ground-based PM2.5 observations across Texas for use in a real-time air quality prediction system. Atmos. Environ. 2005, 39, 7190–7203. [Google Scholar] [CrossRef]
  38. Liu, Y.; Sarnat, J.A.; Kilaru, V.; Jacob, D.J.; Koutrakis, P. Estimating ground-level PM2.5 in the eastern United States using satellite remote sensing. Environ. Sci. Technol. 2005, 39, 3269–3278. [Google Scholar] [CrossRef] [PubMed]
  39. Chu, D.A. Analysis of the relationship between MODIS aerosol optical depth and PM2.5 in the summertime US. Proc. SPIE 2006, 6299, 629903. [Google Scholar]
  40. Engel-Cox, J.A.; Hoff, R.M.; Rogers, R.; Dimmick, F.; Rush, A.C.; Szykman, J.J.; Al-Saadi, J.; Chu, D.A.; Zell, E.R. Integrating lidar and satellite optical depth with ambient monitoring for 3-dimensional particulate characterization. Atmos. Environ. 2006, 40, 8056–8067. [Google Scholar] [CrossRef]
  41. Gupta, P.; Christopher, S.A.; Wang, J.; Gehrig, R.; Lee, Y.; Kumar, N. Satellite remote sensing of particulate matter and air quality assessment over global cities. Atmos. Environ. 2006, 40, 5880–5892. [Google Scholar] [CrossRef]
  42. Koelemeijer, R.B.A.; Homan, C.D.; Matthijsen, J. Comparison of spatial and temporal variations of aerosol optical thickness and particulate matter over Europe. Atmos. Environ. 2006, 40, 5304–5315. [Google Scholar] [CrossRef]
  43. Van Donkelaar, A.; Martin, R.V.; Park, R.J. Estimating ground-level PM2.5 using aerosol optical depth determined from satellite remote sensing. J. Geophys. Res. 2006, 111. [Google Scholar] [CrossRef]
  44. Kumar, N.; Chu, A.; Foster, A. An empirical relationship between PM2.5 and aerosol optical depth in Delhi metropolitan. Atmos. Environ. 2007, 41, 4492–4503. [Google Scholar] [CrossRef] [PubMed]
  45. Wallace, J.; Kanaroglou, P. An investigation of air pollution in southern Ontario, Canada, with MODIS and MISR aerosol data. Int. Geosci. Remote Sens. 2007, 4311–4314. [Google Scholar]
  46. Gupta, P.; Christopher, S.A. Seven year particulate matter air quality assessment from surface and satellite measurements. Atmos. Chem. Phys. 2008, 8, 3311–3324. [Google Scholar] [CrossRef]
  47. Hutchison, K.D.; Faruqui, S.J.; Smith, S. Improving correlations between MODIS aerosol optical thickness and ground-based PM2.5 observations through 3D spatial analyses. Atmos. Environ. 2008, 42, 530–543. [Google Scholar] [CrossRef]
  48. Kumar, N.; Chu, A.; Foster, A. Remote sensing of ambient particles in Delhi and its environs: Estimation and validation. Int. J. Remote Sens. 2008, 29, 3383–3405. [Google Scholar] [CrossRef] [PubMed]
  49. Al-Hamdan, M.Z.; Crosson, W.L.; Limaye, A.S.; Rickman, D.L.; Quattrochi, D.A.; Estes, M.G.; Qualters, J.R.; Sinclair, A.H.; Tolsma, D.D.; Adeniyi, K.A.; et al. Methods for characterizing fine particulate matter using ground observations and remotely sensed data: Potential use for environmental public health surveillance. J. Air Waste Manag. Assoc. 2009, 59, 865–881. [Google Scholar] [CrossRef] [PubMed]
  50. Green, M.; Kondragunta, S.; Ciren, P.; Xu, C. Comparison of GEOS and MODIS aerosol optical depth (AOD) to aerosol robotic network (AERONET) AOD and improve PM2.5 mass at Bondville, Illinois. J. Air Waste Manag. Assoc. 2009, 59, 1082–1091. [Google Scholar] [CrossRef] [PubMed]
  51. Gupta, P.; Christopher, S.A. Particulate matter air quality assessment using integrated surface, satellite, and meteorological products: Multiple regression approach. J. Geophys. Res. 2009, 114. [Google Scholar] [CrossRef]
  52. Gupta, P.; Christopher, S.A. Particulate matter air quality assessment using integrated surface, satellite, and meteorological products: 2. A neural network approach. J. Geophys. Res. 2009, 114. [Google Scholar] [CrossRef]
  53. Paciorek, C.J.; Liu, Y. Limitations of remotely sensed aerosol as a spatial proxy for fine particulate matter. Environ. Health Perspect. 2009, 117, 904–909. [Google Scholar] [CrossRef] [PubMed]
  54. Schaap, M.; Apituley, A.; Timmermans, R.M.A.; Koelemeijer, R.B.A.; 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]
  55. Di Nicolantonio, W.; Cacciari, A. Modis multiannual observations in support of air quality monitoring in northern Italy. Ital. J. Remote Sens. Riv. Ital. Telerilevamento 2010, 43, 97–109. [Google Scholar]
  56. Tian, J.; Chen, D. A semi-empirical model for predicting hourly ground-level fine particulate matter (PM2.5) concentration in southern Ontario from satellite remote sensing and ground-based meteorological measurements. Remote Sens. Environ. 2010, 114, 221–229. [Google Scholar] [CrossRef]
  57. Van Donkelaar, A.; Martin, R.V.; Brauer, M.; Kahn, R.; Levy, R.; Verduzco, C.; Villeneuve, P.J. Global estimates of ambient fine particulate matter concentrations from satellite-based aerosol optical depth: Development and application. Environ. Health Perspect. 2010, 118, 847–855. [Google Scholar] [CrossRef] [PubMed]
  58. Wang, Z.; Chen, L.; Tao, J.; Zhang, Y.; Su, L. Satellite-based estimation of regional particulate matter (PM) in Beijing using vertical-and-RH correcting method. Remote Sens. Environ. 2010, 114, 50–63. [Google Scholar] [CrossRef]
  59. Hu, Z.; Liebens, J.; Rao, K.R. Merging satellite measurement with ground-based air quality monitoring data to assess health effects of fine particulate matter pollution. In Geospatial Analysis of Environmental Health; Maantay, J.A., McLafferty, S., Eds.; Springer: Dordrecht, The Netherlands, 2011; Volume 4, pp. 395–409. [Google Scholar]
  60. Hystad, P.; Setton, E.; Cervantes, A.; Poplawski, K.; Deschenes, S.; Brauer, M.; van Donkelaar, A.; Lamsal, L.; Martin, R.; Jerrett, M.; et al. Creating national air pollution models for population exposure assessment in Canada. Environ. Health Perspect. 2011, 119, 1123–1129. [Google Scholar] [CrossRef] [PubMed]
  61. Wu, Y.; Guo, J.; Zhang, X.; Li, X. Correlation between PM concentrations and aerosol optical depth in eastern China based on BP neural networks. In Proceedings of the 2011 IEEE International Geoscience and Remote Sensing Symposium (IGRASS), Vancouver, BC, Canada, 24–29 July 2011.
  62. Hystad, P.; Demers, P.A.; Johnson, K.C.; Brook, J.; van Donkelaar, A.; Lamsal, L.; Martin, R.; Brauer, M. Spatiotemporal air pollution exposure assessment for a Canadian population-based lung cancer case-control study. Environ. Health 2012, 11, 22. [Google Scholar] [CrossRef] [PubMed]
  63. Lee, S.J.; Serre, M.L.; van Donkelaar, A.; Martin, R.V.; Burnett, R.T.; Jerrett, M. Comparison of geostatistical interpolation and remote sensing techniques for estimating long-term exposure to ambient PM2.5 concentrations across the continental united states. Environ. Health Perspect. 2012, 120, 1727–1732. [Google Scholar] [CrossRef] [PubMed]
  64. Lee, H.J.; Coull, B.A.; Bell, M.L.; Koutrakis, P. Use of satellite-based aerosol optical depth and spatial clustering to predict ambient PM2.5 concentrations. Environ. Res. 2012, 118, 8–15. [Google Scholar] [CrossRef] [PubMed]
  65. Liu, Y.; He, K.; Li, S.; Wang, Z.; Christiani, D.C.; Koutrakis, P. A statistical model to evaluate the effectiveness of PM2.5 emissions control during the Beijing 2008 Olympic Games. Environ. Int. 2012, 44, 100–105. [Google Scholar] [CrossRef] [PubMed]
  66. Mao, L.; Qiu, Y.; Kusano, C.; Xu, X. Predicting regional space-time variation of PM2.5 with land-use regression model and MODIS data. Environ. Sci. Pollut. Res. Int. 2012, 19, 128–138. [Google Scholar] [CrossRef] [PubMed]
  67. Van Donkelaar, A.; Martin, R.V.; Pasch, A.N.; Szykman, J.J.; Zhang, L.; Wang, Y.X.; Chen, D. Improving the accuracy of daily satellite-derived ground-level fine aerosol concentration estimates for north America. Environ. Sci. Technol. 2012, 46, 11971–11978. [Google Scholar] [CrossRef] [PubMed]
  68. Wu, Y.; Guo, J.; Zhang, X.; Tian, X.; Zhang, J.; Wang, Y.; Duan, J.; Li, X. Synergy of satellite and ground based observations in estimation of particulate matter in eastern China. Sci. Total Environ. 2012, 433, 20–30. [Google Scholar] [CrossRef] [PubMed]
  69. Beckerman, B.S.; Jerrett, M.; Martin, R.V.; van Donkelaar, A.; Ross, Z.; Burnett, R.T. Application of the deletion/substitution/addition algorithm to selecting land use regression models for interpolating air pollution measurements in California. Atmos. Environ. 2013, 77, 172–177. [Google Scholar] [CrossRef]
  70. Beckerman, B.S.; Jerrett, M.; Serre, M.; Martin, R.V.; Lee, S.J.; van Donkelaar, A.; Ross, Z.; Su, J.; Burnett, R.T. A hybrid approach to estimating national scale spatiotemporal variability of PM2.5 in the contiguous United States. Environ. Sci. Technol. 2013, 47, 7233–7241. [Google Scholar] [PubMed]
  71. Chudnovsky, A.A.; Kostinski, A.; Lyapustin, A.; Koutrakis, P. Spatial scales of pollution from variable resolution satellite imaging. Environ. Pollut. 2013, 172, 131–138. [Google Scholar] [CrossRef] [PubMed]
  72. Chudnovsky, A.; Lyapustin, A.; Wang, Y.; Schwartz, J.; Koutrakis, P. Analyses of high resolution aerosol data from MODIS satellite: A MAIAC retrieval, southern New England, US. Proc. SPIE 2013, 8795, 8795E-1. [Google Scholar]
  73. Cordero, L.; Wu, Y.; Gross, B.M.; Moshary, F. Assessing satellite AOD based and ARF/CMAQ output PM2.5 estimators. Proc. SPIE 2013, 8723, 872319. [Google Scholar]
  74. Hu, X.; Waller, L.A.; Al-Hamdan, M.Z.; Crosson, W.L.; Estes, M.G., Jr.; Estes, S.M.; Quattrochi, D.A.; Sarnat, J.A.; Liu, Y. Estimating ground-level PM2.5 concentrations in the southeastern U.S. using geographically weighted regression. Environ. Res. 2013, 121, 1–10. [Google Scholar] [CrossRef] [PubMed]
  75. Kumar, N.; Liang, D.; Comellas, A.; Chu, A.D.; Abrams, T. Satellite-based pm concentrations and their application to copd in Cleveland, OH. J. Expo. Sci. Environ. Epidemiol. 2013, 23, 637–646. [Google Scholar] [CrossRef] [PubMed]
  76. Saunders, R.O.; Kahl, J.D.W.; Ghorai, J.K. Improved estimation of PM2.5 using lagrangian satellite-measured aerosol optical depth. Atmos. Environ. 2014, 91, 146–153. [Google Scholar] [CrossRef]
  77. Strawa, A.W.; Chatfield, R.B.; Legg, M.; Scarnato, B.; Esswein, R. Improving retrievals of regional fine particulate matter concentrations from moderate resolution imaging spectroradiometer (MODIS) and ozone monitoring instrument (OMI) multisatellite observations. J. Air Waste Manag. Assoc. 2013, 63, 1434–1446. [Google Scholar] [CrossRef] [PubMed]
  78. Chang, H.H.; Hu, X.; Liu, Y. Calibrating MODIS aerosol optical depth for predicting daily PM2.5 concentrations via statistical downscaling. J. Expo. Sci. Environ. Epidemiol. 2014, 24, 98–404. [Google Scholar] [CrossRef] [PubMed]
  79. Chiu, Y.H.M.; Coull, B.A.; Sternthal, M.J.; Kloog, I.; Schwartz, J.; Cohen, S.; Wright, R.J. Effects of prenatal community violence and ambient air pollution on childhood wheeze in an urban population. J. Allergy Clin. Immunol. 2014, 133, 713–722 e714. [Google Scholar] [CrossRef] [PubMed]
  80. Hu, X.; Waller, L.A.; Lyapustin, A.; Wang, Y.; Al-Hamdan, M.Z.; Crosson, W.L.; Estes, M.G.; Estes, S.M.; Quattrochi, D.A.; Puttaswamy, S.J.; et al. Estimating ground-level PM2.5 concentrations in the southeastern United States using maiac AOD retrievals and a two-stage model. Remote Sens. Environ. 2014, 140, 220–232. [Google Scholar] [CrossRef]
  81. Hu, X.; Waller, L.A.; Lyapustin, A.; Wang, Y.; Liu, Y. 10-year spatial and temporal trends of PM2.5 concentrations in the southeastern us estimated using high-resolution satellite data. Atmos. Chem. Phys. Discuss. 2014, 14, 6301–6314. [Google Scholar] [CrossRef]
  82. Kloog, I.; Chudnovsky, A.A.; Just, A.C.; Nordio, F.; Koutrakis, P.; Coull, B.A.; Lyapustin, A.; Wang, Y.; Schwartz, J. A new hybrid spatio-temporal model for estimating daily multi-year PM2.5 concentrations across northeastern USA using high resolution aerosol optical depth data. Atmos. Environ. 2014, 95, 581–590. [Google Scholar] [CrossRef]
  83. Kloog, I.; Nordio, F.; Zanobetti, A.; Coull, B.A.; Koutrakis, P.; Schwartz, J.D. Short term effects of particle exposure on hospital admissions in the Mid-Atlantic States: A population estimate. PLoS ONE 2014, 9, e88578. [Google Scholar] [CrossRef] [PubMed]
  84. Kim, H.-S.; Chung, Y.-S.; Kim, J.-T. Spatio-temporal variations of optical properties of aerosols in East Asia measured by MODIS and relation to the ground-based mass concentrations observed in central Korea during 2001 similar to 2010. Asia Pac. J. Atmos. Sci. 2014, 50, 191–200. [Google Scholar] [CrossRef]
  85. Lai, H.K.; Tsang, H.; Thach, T.Q.; Wong, C.M. Health impact assessment of exposure to fine particulate matter based on satellite and meteorological information. Environ. Sci. Process Impacts 2014, 16, 239–246. [Google Scholar] [CrossRef] [PubMed]
  86. Lee, H.J.; Kang, C.M.; Coull, B.A.; Bell, M.L.; Koutrakis, P. Assessment of primary and secondary ambient particle trends using satellite aerosol optical depth and ground speciation data in the New England region, United States. Environ. Res. 2014, 133, 103–110. [Google Scholar] [CrossRef] [PubMed]
  87. Ma, Z.; Hu, X.; Huang, L.; Bi, J.; Liu, Y. Estimating ground-level PM2.5 in China using satellite remote sensing. Environ. Sci. Technol. 2014, 48, 7436–7444. [Google Scholar] [CrossRef] [PubMed]
  88. Rush, A.C.; Dougherty, J.J.; Engel-Cox, J.A. Correlating seasonal averaged in-situ monitoring of fine PM with satellite remote sensing data using geographic information system (GIS). Proc. SPIE 2014, 5547, 91–102. [Google Scholar]
  89. Song, W.; Jia, H.; Huang, J.; Zhang, Y. A satellite-based geographically weighted regression model for regional PM2.5 estimation over the Pearl River Delta region in China. Remote Sens. Environ. 2014, 154, 1–7. [Google Scholar] [CrossRef]
  90. Chan, S.H.; van Hee, V.C.; Bergen, S.; Szpiro, A.A.; DeRoo, L.A.; London, S.J.; Marshall, J.D.; Kaufman, J.D.; Sandler, D.P. Long-term air pollution exposure and blood pressure in the sister study. Environ. Health Perspect. 2015, 123, 951–958. [Google Scholar] [CrossRef] [PubMed]
  91. Coker, E.; Ghosh, J.; Jerrett, M.; Gomez-Rubio, V.; Beckerman, B.; Cockburn, M.; Liverani, S.; Su, J.; Li, A.; Kile, M.L.; et al. Modeling spatial effects of PM2.5 on term low birth weight in Los Angeles county. Environ. Res. 2015, 142, 354–364. [Google Scholar] [CrossRef] [PubMed]
  92. Geng, G.; Zhang, Q.; Martin, R.V.; van Donkelaar, A.; Huo, H.; Che, H.; Lin, J.; He, K. Estimating long-term PM2.5 concentrations in China using satellite-based aerosol optical depth and a chemical transport model. Remote Sens. Environ. 2015, 166, 262–270. [Google Scholar] [CrossRef]
  93. Han, Y.; Wu, Y.; Wang, T.; Zhuang, B.; Li, S.; Zhao, K. Impacts of elevated-aerosol-layer and aerosol type on the correlation of AOD and particulate matter with ground-based and satellite measurements in Nanjing, southeast China. Sci. Total Environ. 2015, 532, 195–207. [Google Scholar] [CrossRef] [PubMed]
  94. Just, A.C.; Wright, R.O.; Schwartz, J.; Coull, B.A.; Baccarelli, A.A.; Tellez-Rojo, M.M.; Moody, E.; Wang, Y.; Lyapustin, A.; Kloog, I. Using high-resolution satellite aerosol optical depth to estimate daily PM2.5 geographical distribution in Mexico City. Environ. Sci. Technol. 2015, 49, 8576–8584. [Google Scholar] [CrossRef] [PubMed]
  95. Kloog, I.; Sorek-Hamer, M.; Lyapustin, A.; Coull, B.; Wang, Y.; Just, A.C.; Schwartz, J.; Broday, D.M. Estimating daily PM2.5 and PM10 across the complex geo-climate region of Israel using maiac satellite-based AOD data. Atmos. Environ. 2015, 122, 409–416. [Google Scholar] [CrossRef]
  96. Leon Hsu, H.H.; Chiu, Y.H.M.; Coull, B.A.; Kloog, I.; Schwartz, J.; Lee, A.; Wright, R.O.; Wright, R.J. Prenatal particulate air pollution and asthma onset in urban children. Identifying sensitive windows and sex differences. Am. J. Respir. Crit. Care Med. 2015, 192, 1052–1059. [Google Scholar] [CrossRef] [PubMed]
  97. Lin, C.; Li, Y.; Yuan, Z.; Lau, A.K.H.; Li, C.; Fung, J.C.H. Using satellite remote sensing data to estimate the high-resolution distribution of ground-level PM2.5. Remote Sens. Environ. 2015, 156, 117–128. [Google Scholar] [CrossRef]
  98. McHenry, J.N.; Vukovich, J.M.; Hsu, N.C. Development and implementation of a remote-sensing and in situ data-assimilating version of cmaq for operational PM2.5 forecasting. Part 1: Modis aerosol optical depth (AOD) data-assimilation design and testing. J. Air Waste Manag. Assoc. 2015, 65, 1395–1412. [Google Scholar] [CrossRef] [PubMed]
  99. Nguyen, D.L.; Kim, J.Y.; Ghim, Y.S.; Shim, S.G. Influence of regional biomass burning on the highly elevated organic carbon concentrations observed at Gosan, South Korea during a strong asian dust period. Environ. Sci. Pollut. Res. Int. 2015, 22, 3594–3605. [Google Scholar] [CrossRef] [PubMed]
  100. Song, Y.Z.; Yang, H.L.; Peng, J.H.; Song, Y.R.; Sun, Q.; Li, Y. Estimating PM2.5 concentrations in Xi’an city using a generalized additive model with multi-source monitoring data. PLoS ONE 2015, 10, e0142149. [Google Scholar] [CrossRef] [PubMed]
  101. Van Donkelaar, A.; Martin, R.V.; Spurr, R.J.; Burnett, R.T. High-resolution satellite-derived PM2.5 from optimal estimation and geographically weighted regression over North America. Environ. Sci. Technol. 2015, 49, 10482–10491. [Google Scholar] [CrossRef] [PubMed]
  102. Wong, C.M.; Lai, H.K.; Tsang, H.; Thach, T.Q.; Thomas, G.N.; Lam, K.B.; Chan, K.P.; Yang, L.; Lau, A.K.; Ayres, J.G.; et al. Satellite-based estimates of long-term exposure to fine particles and association with mortality in elderly Hong Kong residents. Environ. Health Perspect. 2015, 123, 1167–1172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Xie, Y.; Wang, Y.; Zhang, K.; Dong, W.; Lv, B.; Bai, Y. Daily estimation of ground-level PM2.5 concentrations over Beijing using 3 km resolution MODIS AOD. Environ. Sci. Technol. 2015, 19, 12280–12288. [Google Scholar] [CrossRef] [PubMed]
  104. Xu, J.; Martin, R.V.; van Donkelaar, A.; Kim, J.; Choi, M.; Zhang, Q.; Geng, G.; Liu, Y.; Ma, Z.; Huang, L.; et al. Estimating ground-level PM2.5 in eastern china using aerosol optical depth determined from the goci satellite instrument. Atmos. Chem. Phys. Discuss. 2015, 15, 17251–17281. [Google Scholar] [CrossRef]
  105. You, W.; Zang, Z.; Pan, X.; Zhang, L.; Chen, D. Estimating PM2.5 in Xi’an, china using aerosol optical depth: A comparison between the MODIS and MISR retrieval models. Sci. Total Environ. 2015, 505, 1156–1165. [Google Scholar] [CrossRef] [PubMed]
  106. Zhang, Y.; Li, Z. Remote sensing of atmospheric fine particulate matter (PM2.5) mass concentration near the ground from satellite observation. Remote Sens. Environ. 2015, 160, 252–262. [Google Scholar] [CrossRef]
  107. Bai, Y.; Wu, L.; Qin, K.; Zhang, Y.; Shen, Y.; Zhou, Y. A geographically and temporally weighted regression model for ground-level PM2.5 estimation from satellite-derived 500 m resolution AOD. Remote Sens. 2016, 8, 262. [Google Scholar] [CrossRef]
  108. Beloconi, A.; Kamarianakis, Y.; Chrysoulakis, N. Estimating urban PM10 and PM2.5 concentrations, based on synergistic meris/aatsr aerosol observations, land cover and morphology data. Remote Sens. Environ. 2016, 172, 148–164. [Google Scholar] [CrossRef]
  109. Crouse, D.L.; Philip, S.; van Donkelaar, A.; Martin, R.V.; Jessiman, B.; Peters, P.A.; Weichenthal, S.; Brook, J.R.; Hubbell, B.; Burnett, R.T. A new method to jointly estimate the mortality risk of long-term exposure to fine particulate matter and its components. Sci. Rep. 2016, 6, 18916. [Google Scholar] [CrossRef] [PubMed]
  110. Di, Q.; Kloog, I.; Koutrakis, P.; Lyapustin, A.; Wang, Y.; Schwartz, J. Assessing PM2.5 exposures with high spatiotemporal resolution across the continental United States. Environ. Sci. Technol. 2016, 50, 4712–4721. [Google Scholar] [CrossRef] [PubMed]
  111. Di, Q.; Koutrakis, P.; Schwartz, J. A hybrid prediction model for PM2.5 mass and components using a chemical transport model and land use regression. Atmos. Environ. 2016, 131, 390–399. [Google Scholar] [CrossRef]
  112. Girguis, M.S.; Strickland, M.J.; Hu, X.; Liu, Y.; Bartell, S.M.; Vieira, V.M. Maternal exposure to traffic-related air pollution and birth defects in Massachusetts. Environ. Res. 2016, 146, 1–9. [Google Scholar] [CrossRef] [PubMed]
  113. He, Q.; Zhou, G.; Geng, F.; Gao, W.; Yu, W. Spatial distribution of aerosol hygroscopicity and its effect on PM2.5 retrieval in east China. Atmos. Res. 2016, 170, 161–167. [Google Scholar] [CrossRef]
  114. Kloog, I. Fine particulate matter (PM2.5) association with peripheral artery disease admissions in northeastern United States. Int. J. Environ. Health Res. 2016, 26, 572–577. [Google Scholar] [CrossRef] [PubMed]
  115. Karimian, H.; Li, Q.; Li, C.; Jin, L.; Fan, J.; Li, Y. An improved method for monitoring fine particulate matter mass concentrations via satellite remote sensing. Aerosol Air Qual. Res. 2016, 16, 1081–1092. [Google Scholar] [CrossRef]
  116. Lee, H.J.; Chatfield, R.B.; Strawa, A.W. Enhancing the applicability of satellite remote sensing for PM2.5 estimation using MODIS deep blue AOD and land use regression in California, United States. Environ. Sci. Technol. 2016, 50, 6546–6555. [Google Scholar] [CrossRef] [PubMed]
  117. Lin, C.; Li, Y.; Lau, A.K.H.; Deng, X.; Tse, T.K.T.; Fung, J.C.H.; Li, C.; Li, Z.; Lu, X.; Zhang, X.; et al. Estimation of long-term population exposure to PM2.5 for dense urban areas using 1-km MODIS data. Remote Sens. Environ. 2016, 179, 13–22. [Google Scholar] [CrossRef]
  118. Lv, B.; Hu, Y.; Chang, H.H.; Russell, A.G.; Bai, Y. Improving the accuracy of daily PM2.5 distributions derived from the fusion of ground-level measurements with aerosol optical depth observations, a case study in north China. Environ. Sci. Technol. 2016, 50, 4752–4759. [Google Scholar] [CrossRef] [PubMed]
  119. Ma, Z.; Hu, X.; Sayer, A.M.; Levy, R.; Zhang, Q.; Xue, Y.; Tong, S.; Bi, J.; Huang, L.; Liu, Y. Satellite-based spatiotemporal trends in PM2.5 concentrations: China, 2004–2013. Environ. Health Perspect. 2016, 124, 184–192. [Google Scholar] [CrossRef] [PubMed]
  120. Shi, L.; Zanobetti, A.; Kloog, I.; Coull, B.A.; Koutrakis, P.; Melly, S.J.; Schwartz, J.D. Low-concentration pm and mortality: Estimating acute and chronic effects in a population-based study. Environ. Health Perspect. 2016, 124, 46–52. [Google Scholar] [PubMed]
  121. Strickland, M.J.; Hao, H.; Hu, X.; Chang, H.H.; Darrow, L.A.; Liu, Y. Pediatric emergency visits and short-term changes in pm concentrations in the U.S. State of Georgia. Environ. Health Perspect. 2016, 124, 6900–6696. [Google Scholar]
  122. Stieb, D.M.; Chen, L.; Beckerman, B.S.; Jerrett, M.; Crouse, D.L.; Omariba, D.W.; Peters, P.A.; van Donkelaar, A.; Martin, R.V.; Burnett, R.T.; et al. Associations of pregnancy outcomes and PM in a national Canadian study. Environ. Health Perspect. 2016, 124, 243–249. [Google Scholar] [PubMed]
  123. Van Donkelaar, A.; Martin, R.V.; Brauer, M.; Hsu, N.C.; Kahn, R.A.; Levy, R.C.; Lyapustin, A.; Sayer, A.M.; Winker, D.M. Global estimates of fine particulate matter using a combined geophysical-statistical method with information from satellites, models, and monitors. Environ. Sci. Technol. 2016, 50, 3762–3772. [Google Scholar] [CrossRef] [PubMed]
  124. Wang, B.; Chen, Z. High-resolution satellite-based analysis of ground-level PM2.5 for the city of Montreal. Sci. Total Environ. 2016, 541, 1059–1069. [Google Scholar] [CrossRef] [PubMed]
  125. You, W.; Zang, Z.; Zhang, L.; Li, Y.; Pan, X.; Wang, W. National-scale estimates of ground-level PM2.5 concentration in china using geographically weighted regression based on 3 km resolution MODIS AOD. Remote Sens. 2016, 8, 184. [Google Scholar] [CrossRef]
  126. You, W.; Zang, Z.; Zhang, L.; Li, Y.; Wang, W. Estimating national-scale ground-level PM2.5 concentration in china using geographically weighted regression based on MODIS and MISR AOD. Environ. Sci. Pollut. Res. Int. 2016, 23, 8327–8338. [Google Scholar] [CrossRef] [PubMed]
  127. Zheng, Y.; Zhang, Q.; Liu, Y.; Geng, G.; He, K. Estimating ground-level PM2.5 concentrations over three megalopolises in china using satellite-derived aerosol optical depth measurements. Atmos. Environ. 2016, 124, 232–242. [Google Scholar] [CrossRef]
  128. Zou, B. High-resolution satellite mapping of fine particulates based on geographically weighted regression. IEEE Geosci. Remote Sens. Lett. 2016, 13, 495–499. [Google Scholar] [CrossRef]
  129. Guo, H.; Cheng, T.; Gu, X.; Gu, X.; Chen, H.; Wang, Y.; Zheng, F.; Xiang, K. Comparison of four ground-level PM2.5 estimation models using parasol aerosol optical depth data from China. Int. J. Environ. Res. Public Health 2016, 13, 180. [Google Scholar] [CrossRef] [PubMed]
  130. Kloog, I.; Coull, B.A.; Zanobetti, A.; Koutrakis, P.; Schwartz, J.D. Acute and chronic effects of particles on hospital admissions in New-England. PLoS ONE 2012, 7, e34664. [Google Scholar] [CrossRef] [PubMed]
  131. Kloog, I.; Ridgway, B.; Koutrakis, P.; Coull, B.A.; Schwartz, J.D. Long- and short-term exposure to PM2.5 and mortality: Using novel exposure models. Epidemiology 2013, 24, 555–561. [Google Scholar] [CrossRef] [PubMed]
  132. Madrigano, J.; Kloog, I.; Goldberg, R.; Coull, B.A.; Mittleman, M.A.; Schwartz, J. Long-term exposure to PM2.5 and incidence of acute myocardial infarction. Environ. Health Perspect. 2013, 121, 192–196. [Google Scholar] [PubMed]
  133. Lakshmanan, A.; hiu, Y.H.; Coull, B.A.; Just, A.C.; Maxwell, S.L.; Schwartz, J.; Gryparis, A.; Kloog, I.; Wright, R.J.; Wright, R.O. Associations between prenatal traffic-related air pollution exposure and birth weight: Modification by sex and maternal pre-pregnancy body mass index. Environ. Res. 2015, 137, 268–277. [Google Scholar] [CrossRef] [PubMed]
  134. Al-Hamdan, M.Z.; Crosson, W.L.; Economou, S.A.; Estes, M.G., Jr.; Estes, S.M.; Hemmings, S.N.; Kent, S.T.; Puckett, M.; Quattrochi, D.A.; Rickman, D.L.; et al. Environmental public health applications using remotely sensed data. Geocarto Int. 2014, 29, 85–98. [Google Scholar] [CrossRef] [PubMed]
  135. Alexeeff, S.E.; Schwartz, J.; Kloog, I.; Chudnovsky, A.; Koutrakis, P.; Coull, B.A. Consequences of kriging and land use regression for PM2.5 predictions in epidemiologic analyses: Insights into spatial variability using high-resolution satellite data. J. Expo. Sci. Environ. Epidemiol. 2015, 25, 138–144. [Google Scholar] [CrossRef] [PubMed]
  136. Ma, Z.; Liu, Y.; Zhao, Q.; Liu, M.; Zhou, Y.; Bi, J. Satellite-derived high resolution PM2.5 concentrations in yangtze river delta region of China using improved linear mixed effects model. Atmos. Environ. 2016, 133, 156–164. [Google Scholar] [CrossRef]
  137. Kloog, I.; Melly, S.J.; Ridgway, W.L.; Coull, B.A.; Schwartz, J. Using new satellite based exposure methods to study the association between pregnancy PM2.5 exposure, premature birth and birth weight in massachusetts. Environ. Health. 2012, 11, 40. [Google Scholar] [CrossRef] [PubMed]
  138. Liu, Y. Validation of multiangle imaging spectroradiometer (MISR) aerosol optical thickness measurements using aerosol robotic network (aeronet) observations over the contiguous United States. J. Geophys. Res. 2004, 109. [Google Scholar] [CrossRef]
  139. Konkel, L. The view from afar satellite-derived estimates of global PM2.5. Environ. Health Perspect. 2015, 123, A43. [Google Scholar] [CrossRef] [PubMed]
  140. Spivey, A. Keeping an eye on PM2.5: Satellite data reveal global picture of particulate pollution. Environ. Health Perspect. 2010, 118, A259. [Google Scholar] [CrossRef] [PubMed]
  141. Crouse, D.L.; Peters, P.A.; van Donkelaar, A.; Goldberg, M.S.; Villeneuve, P.J.; Brion, O.; Khan, S.; Atari, D.O.; Jerrett, M.; Pope, C.A.; et al. Risk of non-accidental and cardiovascular mortality in relation to long-term exposure to low concentrations of fine particulate matter: A Canadian national-level cohort study. Environ. Health Perspect. 2012, 120, 708–714. [Google Scholar] [CrossRef] [PubMed]
  142. Chen, B.; Sverdlik, L.; Imashev, S.; Solomon, P.; Lantz, J.; Schauer, J.; Shafer, M.; Artamonova, M.; Carmichael, G. Empirical relationship between particulate matter and aerosol optical depth over northern Tien-Shan, central Asia. Air Qual. Atmos. Health. 2013, 6, 385–396. [Google Scholar] [CrossRef]
  143. Villeneuve, P.J.; Goldberg, M.S.; Burnett, R.T.; van Donkelaar, A.; Chen, H.; Martin, R.V. Associations between cigarette smoking, obesity, sociodemographic characteristics and remote-sensing-derived estimates of ambient PM2.5: Results from a Canadian population-based survey. Occup. Environ. Med. 2011, 68, 920–927. [Google Scholar] [CrossRef] [PubMed]
  144. To, T.; Zhu, J.; Villeneuve, P.J.; Simatovic, J.; Feldman, L.; Gao, C.; Williams, D.; Chen, H.; Weichenthal, S.; Wall, C.; et al. Chronic disease prevalence in women and air pollution—A 30-year longitudinal cohort study. Environ. Int. 2015, 80, 26–32. [Google Scholar] [CrossRef] [PubMed]
  145. Brook, R.D.; Cakmak, S.; Turner, M.C.; Brook, J.R.; Crouse, D.L.; Peters, P.A.; van Donkelaar, A.; Villeneuve, P.J.; Brion, O.; Jerrett, M.; et al. Long-term fine particulate matter exposure and mortality from diabetes in Canada. Diabetes Care 2013, 36, 3313–3320. [Google Scholar] [CrossRef] [PubMed]
  146. Jerrett, M.; Burnett, R.T.; Beckerman, B.S.; Turner, M.C.; Krewski, D.; Thurston, G.; Martin, R.V.; van Donkelaar, A.; Hughes, E.; Shi, Y.; et al. Spatial analysis of air pollution and mortality in California. Am. J. Respir. Crit. Care Med. 2013, 188, 593–599. [Google Scholar] [CrossRef] [PubMed]
  147. Van Donkelaar, A.; Martin, R.V.; Spurr, R.J.D.; Drury, E.; Remer, L.A.; Levy, R.C.; Wang, J. Optimal estimation for global ground-level fine particulate matter concentrations. J. Geophys. Res. Atmos. 2013, 118, 5621–5636. [Google Scholar] [CrossRef]
  148. 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] [PubMed]
  149. Lin, G.; Fu, J.; Jiang, D.; Hu, W.; Dong, D.; Huang, Y.; Zhao, M. Spatio-temporal variation of PM2.5 concentrations and their relationship with geographic and socioeconomic factors in China. Int. J. Environ. Res. Public Health 2014, 11, 173–186. [Google Scholar] [CrossRef] [PubMed]
  150. Brauer, M.; Amann, M.; Burnett, R.T.; Cohen, A.; Dentener, F.; Ezzati, M.; Henderson, S.B.; Krzyzanowski, M.; Martin, R.V.; van Dingenen, R.; et al. Exposure assessment for estimation of the global burden of disease attributable to outdoor air pollution. Environ. Sci. Technol. 2012, 46, 652–660. [Google Scholar] [CrossRef] [PubMed]
  151. Brunsdon, C.; Fotheringham, A.S.; Charlton, M.E. Geographically weighted regression: A method for exploring spatial nonstationarity. Geogr. Anal. 1996, 28, 281–298. [Google Scholar] [CrossRef]
  152. Fotheringham, A.S.; Charlton, M.E.; Brunsdon, C. Geographically weighted regression: A natural evolution of the expansion method for spatial data analysis. Environ. Plan. A 1998, 30, 1905–1927. [Google Scholar] [CrossRef]
  153. Fotheringham, A.S.; Brunsdon, C.; Charlton, M. Geographically weighted regression: The analysis of spatically varying relationships. In Geographical Analysis; O’Sullivan, D., Ed.; Wiley: NewYork, USA, 2003; Volume 35, pp. 272–275. [Google Scholar]
Figure 1. Flow chart of study selection.
Figure 1. Flow chart of study selection.
Atmosphere 07 00129 g001
Figure 2. The frequency distribution of seven models.
Figure 2. The frequency distribution of seven models.
Atmosphere 07 00129 g002
Figure 3. Constituent ratio of seven models.
Figure 3. Constituent ratio of seven models.
Atmosphere 07 00129 g003
Table 1. Characteristics of included studies.
Table 1. Characteristics of included studies.
Author (Published Year)Study AreaStudy PeriodSource of AODRetrieved ModelR2 of Model (CV-R2)
Wang et al. (2003) [16]U.S.2002MODISLC0.960 a (Nss = 1, Nms = 7)
Engel-Cox et al. (2004) [31]U.S.2002MODISLC0.185 a,b
Liu et al. (2004) [18]U.S.2001MISRCTM0.656 a,b (Yearly, Nms = 1268)
Hutchison et al. (2005) [37]U.S.2003–2004MODISLC0.160~0.250 a,b (Nms = 51)
Liu et al. (2005) [38]U.S.2001MISRMLR0.430 a,b (Nms = 346)
Chu et al. (2006) [39]U.S.2002MODISMLR0.723 a (New York), 0.757 a (Chicago), 0.774 a (Houston) (Nms = 350 for U.S.)
Engel-Cox et al. (2006) [40]U.S.2004MODISMLR0.423 a
Gupta et al. (2006) [41]Global2000–2002MODISMLR0.960 a,b (Nss = 26, Nms = 113)
Kacenelenbogen et al. (2006) [26]France2003POLDERMLR0.490 a,b (when the matched data is 78),
0.310 a,b (Nms = 28, when the matched data is 1974)
Koelemeijer et al. (2006) [42]Europe2003MODISLC0.360 a,b (Nms = 88)
van Donkelaar et al. (2006) [43]Global2000–2001MODIS, MISRCTM0.476 a,b (MODIS, Nms = 199),
0.325 a,b (MISR, Nms = 199)
Kumar et al. (2007) [44]India2003MODISMLR0.700 a,b (Point/disaggregate-level analysis, Nms = 113),
0.610 a,b (Aggregate/pixel-level analysis, Nms = 113)
Liu et al. (2007) [13]U.S.2005MISRCTMEastern: 0.560 a,b (with fractional AOD, Nms = 130), 0.420 a,b (with total AOD, Nms = 130)
Western: 0.570 a,b (with fractional AOD, Nms = 130), 0.210 a,b (with total AOD, Nms = 130)
Liu et al. (2007) [20]U.S.2003MODIS, MISRGLM0.510 a,b (MODIS, St. Louis and its surrounding counties, Nms = 22),
0.620 a,b (MISR, St. Louis and its surrounding counties, Nms = 22)
Wallace et al. (2007) [45]Canada2015MODISMLR0.760 b (Nms = 34)
Gupta et al. (2008) [46]U.S.2000–2006MODISMLR0.520 a,b (Daily, Nms = 14),
0.620 a,b (Hourly, Nms = 14)
Gupta et al. (2008) [11]U.S.2000–2005MODISMLR0.270 a,b (Nms = 38)
Hutchison et al. (2008) [47]U.S.2003, 2004MODISMLR0.221 a,b (20 August–15 September, Hourly, Houston-Beaumont-Galveston area),
0.960 a,b (6–7 September, Hourly, Houston-Beaumont-Galveston area)
Kumar et al. (2008) [48]India2003MODISMLR0.700 a,b (Point/disaggregate-level analysis, Delhi and its environs, Nms = 113),
0.610 a,b (Aggregate/pixel-level analysis, Delhi and its environs, Nms = 113)
Paciorek et al. (2008) [24]U.S.2004MODIS, MISR, GEOSGAM0.360 a,b
Al-Hamdan et al. (2009) [49]U.S.2000–2003MODISMLR0.661~0.706 a,b (MODIS), 0.874 a,b (B-Spline, merged AQS/MODIS),
0.949 a,b (IDW, merged AQS/MODIS)
Green et al. (2009) [50]U.S.2003–2007GEOS, MODISMLR0.480 a (GEOS, Nss = 1), 0.740 a (MODIS, Nss = 1)
Gupta et al. (2009) [51]U.S.2004–2006MODISMLR0.365 a,b (TVM, Nms = 85), 0.466 a,b (MVM, Nms = 85)
Gupta et al. (2009) [52]U.S.2004–2006MODISANN0.608 a,b (Nms = 85)
Hu et al. (2009) [32]U.S.2003–2004MODISGWR, LC0.449 a,b (LC, East), 0.048 a,b (LC, West); 0~0.580 a,b (GWR, Nms = 877)
Liu et al. (2009) [23]U.S.2003–2005GEOSGAM0.790 a,b (Adjusted, Nms = 32), 0.480 a,b (Unadjusted, Nms = 32);
0.780 *,a,b (Adjusted, Nms = 32), 0.460 *,a,b (Unadjusted, Nms = 32)
Paciorek et al. (2009) [53]U.S.2004MODIS, MISR, GEOSGAM0.573 a,b (MODIS, Yearly), 0.572 a,b (GEOS, Yearly);
0.825 a,b (MODIS, Monthly), 0.825 a,b (GEOS, Monthly)
Schaap et al. (2009) [54]Netherlands2006–2007MODISMLR0.518 a,b
Zhang et al. (2009) [21]U.S.2005–2006MODISMLR0.600 a,b (Southeast U.S.), 0.200 a,b (Southwest U.S.), (Nms = 521 for U.S.)
Di Nicolantonio et al. (2010) [55]Italy2007MODISCTM0.680 a,b (Terra MODIS, Nms = 23), 0.590 a,b (Aqua MODIS, Nms = 23),
0.700 a,b (Terra and Aqua MODIS, Nms = 23)
Leon et al. (2010) [25]Europe, Africa2006–2008POLDERMLR0.250 a,b (Nms = 28)
Tian et al. (2010) [56]Canada2004MODISSemi-empirical model0.650 a,b (Hourly, Nms = 30)
van Donkelaar et al. (2010) [57]Global2001–2006MODIS, MISRCTM0.593 a,b (North America, Nms = 1057),
0.689 a,b (Elsewhere, Nms = 244)
Wang et al. (2010) [58]China2007–2008MODISLC0.470 a (Nss = 1, Nms = 20)
Hu et al. (2011) [59]U.S.2003–2004MODISGWR, LC0~1 a,b (GWR, Nms = 877), 0.449 a,b (LC, Nms = 877)
Hystad et al. (2011) [60]Canada2006MODIS, MISRLUR0.460 *,a,b (Nms = 177)
Kloog et al. (2011) [33]U.S.2000–2008MODISMEM0.830 *,a,b (with available AOD, Nms = 78),
0.810 *,a,b (without available AOD, Nms = 78)
Lee et al. (2011) [19]U.S.2003MODISMEM0.970 a,b (Nms = 26), 0.920 *,a,b (Nms = 26)
Wu et al. (2011) [61]China2007–2008MODISANN0.030 a,b (Hourly in summer, Nms = 10),
0.580 a,b (Hourly in winter, Nms = 10)
Chudnovsky et al. (2012) [35]U.S.2003GEOSMEM0.970 a,b (Nms = 26), 0.920 *,a,b (Nms = 26)
Hystad et al. (2012) [62]Canada1975–1994MODIS, MISRCTM0.670 a,b (Nms = 25)
Kloog et al. (2012) [34]U.S.2000–2008MODISMEM0.850 *,a,b (Nss = 8, Nms = 161)
Lee et al. (2012) [63]U.S.2001–2006MODIS, MISRCTM0.200~0.820 a,b
Lee et al. (2012) [64]U.S.2000–2008MODISMEM0.930 a,b (MEM for available AOD, Nms = 69),
0.880 *,a,b (MEM for available AOD, Nms = 69)
Liu et al. (2012) [65]China2008MODISGAM0.563 a (Adjusted, Nss = 1, Nms = 3); 0.757 a (Unadjusted, Nss = 1, Nms = 3);
0.372 *,a (Adjusted, Nss = 1, Nms = 3), 0.608 *,a (Unadjusted, Nss = 1, Nms = 3)
Mao et al. (2012) [66]U.S.2005MODISLUR0.648 a,b (Unadjusted, Nms = 34), 0.626 a,b (Adjusted, Nms = 34), 0.58 *,a,b (Nms = 34)
van Donkelaar et al. (2012) [67]U.S.2004–2009MODIS, MISRCTM0.689 b (for day of June 27, 2005. Nms = 1482)
Wu et al. (2012) [68]China2007–2008MODISANN0.430 a,b (Nms = 7)
Beckerman et al. (2013) [69]U.S.2001–2006-LUR0.650 *,a,b (Monthly, Nms = 4119)
Beckerman et al. (2013) [70]U.S.1991–2008GEOSLUR0.630 *,a,b (LUR, Nms = 1464), 0.790 *,a,b (LUR and BMEM, Nms = 1464)
Chudnovsky et al. (2013) [71]U.S.2003MODISLC0.470 a (New England), 0.620 a (Boston), Nms = 26 for U.S.
Chudnovsky et al. (2013) [72]U.S.2002–2008MODISMEM0.500 *,a,b (New England), 0.860 *,a,b (Boston), Nms = 26 for U.S.
Cordero et al. (2013) [73]U.S.2005–2006MODIS, GEOSMLR0.860 a (Urban areas in summer, Nms = 39)
Hu et al. (2013) [74]U.S.2003MODISGWR0.600 a,b (NARR, Nms = 119), 0.610 a,b (NLDAS, Nms = 119),
0.672 *,a,b (NARR, Nms = 119), 0.706 *,a,b (NLDAS, Nms = 119)
Kumar et al. (2013) [75]U.S.2000–2009MODISMLR0~1 a,b (Nms = 5)
Saunders et al. (2013) [76]U.S.2003–2007MODISMLR0.760 a,b (Winter)
Strawa et al. (2013) [77]U.S.2004–2008MODISGAM0.770 a,b
Tao et al. (2013) [17]China2007–2008MODISMLR0.610 a,b (Beijing and its surrounding regions, Nms = 17)
Chang et al. (2014) [78]U.S.2003–2005MODISLUR0.780 *,a,b (Nms = 85)
Chiu et al. (2014) [79]U.S.2002–2009MODISMEM0.830 *,a,b (with available AOD, Nms = 78);
0.810 *,a,b (without available AOD, Nms = 78)
Hu et al. (2014) [80]U.S.2003MODISTSM0.830 a,b,0.670 *,a,b
Hu et al. (2014) [81]U.S.2001–2010MODIS, MISRTSM0.710~0.850 a,b (for year 2001–2010),
0.62~0.78 a,b (for year 2001–2010)
Kloog et al. (2014) [82]U.S.2003–2011MODISMEM0.880 *,a,b (Nms = 161)
Kloog et al. (2014) [83]U.S.2000–2006MODISMEM0.810 *,a,b (Nms = 161)
Kim et al. (2014) [84]Korea2001–2010MODISCTM0.440 *,a,b (for PM2.5 sulphate), 0.370 *,a,b (for PM2.5 dust),
0.230 *,a,b (for PM2.5 smoke)
Lai et al. (2014) [85]Global2012MODISMLR0.850 a,b (The best, Nms = 31)
Lary et al. (2014) [28]Global1997–2014Sea WIFS, MODISMachine-learning regression0.920 a,b (N = 8329)
Lee et al. (2014) [86]U.S.2000–2008MODISMEM0.890 a,b (for retrieval days, Nms = 69),
0.860 *,a,b (for retrieval days, Nms = 69),
0.790 *,a,b (for non-retrieval days, Nms = 69)
Ma et al. (2014) [87]China2012–2013MODIS, MISRGWR0.710 a,b (Nss = 113, Nms = 835), 0.640 *,a,b (Nss = 113, Nms = 835)
Rush et al. (2014) [88]U.S.2001MODISKriging0.815 b (Northeast summer);
0.800 b (Industrial Midwest summer)
Song et al. (2014) [89]China2012–2013MODISGWR0.738 a,b (PRD, Nms = 37)
Toth et al. (2014) [30]U.S.2008–2009MODIS, MISR, CALIOPLC0.130 a,b (Aqua MODIS, Hourly, Nms = 102),
0.090 a,b (Terra MODIS, Hourly, Nms = 102),
0.090 a,b (MISR, Hourly, Nms = 102);
0.040 a,b (Aqua MODIS, Daily, Nms = 991),
0.063 a,b (Terra MODIS, Daily, Nms = 991),
0.063 a,b (MISR, Daily, Nms = 991)
Chan et al. (2015) [90]U.S.2003–2009MODISKriging0.880 *,a,b
Coker et al. (2015) [91]U.S.1995–2006-LUR0.650 *,a,b
Geng et al. (2015) [92]China2006–2012MODIS, MISRCTM0.548 a,b (Nms = 46)
Han et al. (2015) [93]China2011MODISMLR0.624 a (All dust data but filter out aloft-dust-layer, Nss = 1);
0.548 a (All non-dust data, Nss = 1)
Just et al. (2015) [94]Mexico2004–2014MODISMEM0.724 *,a (Nss = 1, Nms = 12)
Kloog et al. (2015) [95]Israel2003–2013MODISMEM0.720 *,a,b (Nms = 45)
Leon Hsu et al. (2015) [96]U.S.2002–2009MISRMEM0.830 *,a,b (with available AOD, Nms = 78),
0.810 *,a,b (without available AOD, Nms = 78)
Lee et al. (2015) [12]U.S.2007–2011MODISMEM0.770 *,a,b, 0.810 *,a,b, 0.700 *,a,b for region 1, 2, 3 (Nms = 277)
Lee et al. (2015) [7]U.S.2003–2011MODISMEM0.770 *,a,b, 0.810 *,a,b, 0.700 *,a,b for region 1, 2, 3 (Nms = 257)
Li et al. (2015) [29]U.S.2005–2010MODIS, MISR, SeaWiFS, OMICMCA, MCACMCA: 0.600 a,b (MODIS/MISRR/SeaWiFS/OMI, Nms = 98),
0.792 a,b (for year between 2005 and 2010, Nms = 198);
MCA: 0.828 a,b (for year between 2005 and 2010, Nms = 98)
Lin et al. (2015) [97]China2013MODISSemi-empirical model0.810 a,b (Nms = 565, Yearly), 0.578a,b (Nms = 565, Monthly)
McHenry et al. (2015) [98]U.S.2002MODISCMAQ0.468 a,b (yearly)
Nguyen et al. (2015) [99]Vietnam2011–2012MODISSVR, MLR0.352 a,b (SVR), 0.358 a,b (MLR)
Song et al. (2015) [100]China2013MODISGAM0.691 a (Nss = 1, Nms = 13)
van Donkelaar et al. (2015) [101]U.S.2004–2008MODISCTM0.620 a,b (Unadjusted, Nms = 1253), 0.820 a,b (Adjusted, Nms = 1253),
0.780 *,a,b (Nms = 1253)
van Donkelaar et al. (2015) [27]Global1998–2012MODIS, MISR SeaWiFSCTM0.656 a,b (North America and Europe, Nms = 210)
Wong et al. (2015) [102]China2000–2011-SEC0.360
Xie et al. (2015) [103]China2013–2014MODISMEM0.810~0.830 a (various between districts, Nss = 1, Nms = 35),
0.750~0.790 *,a (various between districts, Nss = 1, Nms = 35)
Xu et al. (2015) [104]China2013GOCICTM0.656 a,b (Yearly, Nms = 494)
You et al. (2015) [105]China2013MODIS, MISRNonlinear regression model0.670 a (MODIS, Nss = 1, Nms = 13),
0.720 a (MISR, Nss = 1, Nms = 13)
Zhang et al. (2015) [106]China2013MODISMLR0.462 a (Hourly, Nss = 1, Nms = 15)
Bai et al. (2016) [107]China2015MODISGTWR, OLS, GWR, TWR0.960 a,b (GTWR, Nms = 37), 0.870 *,a,b (GTWR, Nms = 37);
0.350 a,b (OLS, Nms = 37), 0.410 a,b (OLS, Nms = 37);
0.590 a,b (GWR, Nms = 37), 0.600 a,b (GWR, Nms = 37);
0.630 a,b (TWR, Nms = 37), 0.680 a,b (TWR, Nms = 37)
Beloconi et al. (2016) [108]UK2002–2012MODISKriging, MEM0.040 *,a (Kriging, Nss = 1),0.846 *,a (MEM, Nss = 1)
Crouse et al. (2016) [109]Canada2001–2010MODIS, MISR, SeaWiFSCTM0.578 a,b
Di et al. (2016) [110]U.S.2000–2012MODISANN0.840 * a,b (Nms = 1928)
Di et al. (2016) [111]U.S.2001–2010-ANN0.850 ** a,b (Nms = 154)
Girguis et al. (2016) [112]U.S.2001–2008MODISMEM0.780~0.880 *,a,b (for year 2001–2008, Nms = 35)
He et al. (2016) [113]China2014–2015MODISLC0.723 a,b (Nss = 6, Nms = 82)
Kloog et al. (2016) [114]U.S.2000–2008MODIS, MISRMEM0.820 *,a,b
Karimian et al. (2016) [115]China2013MODISImproved LC0.500 a (Terra MODIS, Nss = 1, Nms = 8),
0.566 a (Aqua MODIS, Nss = 1, Nms = 8)
Lee et al. (2016) [116]U.S.2006–2012MODISMEM0.666 *,a,b (Nms = 87)
Lin et al. (2016) [117]China2000–2014MODISLC0.672 a,b (Monthly, 2000–2014, Nms = 3094),
0.608 a,b (Yearly, 2013, Nms = 76), 0.548 (Yearly, 2014, Nms = 86)
Lv et al. (2016) [118]China2014MODISBayesian model0.780 *,a,b (Nss = 53, Nms = 298)
Ma et al. (2016) [87]China2013MODISImproved MEM0.725 *,a,b (Nested MEM, Nss = 5, Nms = 115),
0.724 *,a,b (Non-nested MEM, Nss = 5, Nms = 115);
0.486 **,a,b (Nested MEM, Nss = 5, Nms = 115),
0.230 **,a,b (Non-nested MEM, Nss = 5, Nms = 115)
Ma et al. (2016) [119]China2004–2013MODISTSM0.790 *,a,b (Nss = 205, Nms = 1185)
Shi et al. (2016) [120]U.S.2003–2008MODISMEM0.870 *,a,b
Strickland et al. (2016) [121]U.S.2002–2010MODISTSM0.710~0.85 a,b (Yearly)
Stieb et al. (2016) [122]Canada1999–2008MODISLUR0.590 *,a,b (Nms = 241)
van Donkelaar et al. (2016) [123]Global1998–2014MODIS, MISR, SeaWiFSCTM and GWR0.810 *,a,b
Wang et al. (2016) [124]Canada2009MODISCTM0.860 a (Daily, Nss = 1, Nms = 10), 0.930 a (Monthly, Nss = 1, Nms = 10)
You et al. (2016) [125]China2014MODISGWR0.810 a,b (Nms = 943), 0.790 *,a,b (Nms = 943)
You et al. (2016) [126]China2014MODIS, MISRGWR0.760 *,a,b (MODIS, Nms = 943), 0.810 *,a,b (MISR, Nms = 943)
Zheng et al. (2016) [127]China2013MODISMEM0.770 *,a,b (BTH, Nss = 3, Nms = 66), 0.800 *,a,b (YRD, Nss = 15, Nms = 56),
0.800 *,a,b (PRD, Nss = 11, Nms = 55)
Zou et al. (2016) [128]China2013MODISGWR, OLS0.750 a,b (GWR, Nss = 3, Nms = 52), 0.530 a,b (OLS, Nss = 3, Nms = 52)
* Sample-based CV-R2: Sample-based Cross Validated-coefficient of determination; ** DOY-based CV-R2: Day-of Year-based Cross Validated-coefficient of determination; a,b denotes Temporal and Spatial of R2, respectively; R2 denotes daily PM2.5 expect for note with monthly and yearly; Nms denotes number of PM2.5 monitoring site; Nss denotes number of study site at city level. The list of abbreviations: (1) Satellite Sensors: MODIS, Moderate Resolution Imaging Spectrometer; MISR, Multi-Angle Imaging Spectrometer; GEOS, Geostationary Operational Environment Satellite; SeaWiFS, Sea-viewing Wide Field-of-view Sensor; POLDER, Polarization of Earth’s Reflectance and Directionality; CALIOP, Cloud-Aerosol Lidar with Orthogonal Polarization; GOCI, Geostationary Ocean Color Imager; OMI, Ozone Monitoring Instrument; (2) Derived models: LC, Linear Correlations; MLR, Multiple Linear Regression; LUR, Land Use Regression; GAM, Generalized Additive Model; MEM, Mixed-Effect Model; CTM, Chemical Transport Model; GLM, General Linear regression Model; ANN, Artificial Neural Networks; TSM, Two-Stage Model; SVR, Support Vector Regression; GTWR, Geographically and Temporally Weighted Regression; TWR, Temporally Weighted Regression; TVM, Two-Variate Method; MVM, Multivariate Method; OLS, Ordinary Least Squares model; SEC, Surface Extinction Coefficient; Nested MEM, Nested Mixed Effects Model; Non-nested MEM, Non-nested Mixed Effects Model; DSA, Deletion/substitution/addition; BMEM, Bayesian Maximum Entropy method; MCA, Maximum Covariance Analysis; CMCA, Combined Maximum Covariance Analysis (3) R2 of model, Coefficient of determination of model: NARR, North American Regional Reanalysis; NLDAS, North American Land Data Assimilation System; BTH, Beijing-Tianjin-Hebei region; YRD, Yangtze River Delta region; PRD, Pearl River Delta region.

Share and Cite

MDPI and ACS Style

Chu, Y.; Liu, Y.; Li, X.; Liu, Z.; Lu, H.; Lu, Y.; Mao, Z.; Chen, X.; Li, N.; Ren, M.; et al. A Review on Predicting Ground PM2.5 Concentration Using Satellite Aerosol Optical Depth. Atmosphere 2016, 7, 129. https://doi.org/10.3390/atmos7100129

AMA Style

Chu Y, Liu Y, Li X, Liu Z, Lu H, Lu Y, Mao Z, Chen X, Li N, Ren M, et al. A Review on Predicting Ground PM2.5 Concentration Using Satellite Aerosol Optical Depth. Atmosphere. 2016; 7(10):129. https://doi.org/10.3390/atmos7100129

Chicago/Turabian Style

Chu, Yuanyuan, Yisi Liu, Xiangyu Li, Zhiyong Liu, Hanson Lu, Yuanan Lu, Zongfu Mao, Xi Chen, Na Li, Meng Ren, and et al. 2016. "A Review on Predicting Ground PM2.5 Concentration Using Satellite Aerosol Optical Depth" Atmosphere 7, no. 10: 129. https://doi.org/10.3390/atmos7100129

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