Next Article in Journal
An Accurate TLS and UAV Image Point Clouds Registration Method for Deformation Detection of Chaotic Hillside Areas
Previous Article in Journal
Preliminary Assessment of Turbidity and Chlorophyll Impact on Bathymetry Derived from Sentinel-2A and Sentinel-3A Satellites in South Florida
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Estimation of PMx Concentrations from Landsat 8 OLI Images Based on a Multilayer Perceptron Neural Network

1
School of Human Settlements and Civil Engineering, Xi’An Jiaotong University, Xi’an 710049, China
2
Remote Sensing Technology Institute (IMF), German Aerospace Center (DLR), 82234 Wessling, Germany
3
Signal Processing in Earth Observation (SiPEO), Technical University of Munich (TUM), 80333 Munich, Germany
*
Author to whom correspondence should be addressed.
Remote Sens. 2019, 11(6), 646; https://doi.org/10.3390/rs11060646
Submission received: 16 February 2019 / Revised: 25 February 2019 / Accepted: 11 March 2019 / Published: 16 March 2019
(This article belongs to the Section Atmospheric Remote Sensing)

Abstract

:
The estimation of PMx (incl. PM10 and PM2.5) concentrations using satellite observations is of great significance for detecting environmental issues in many urban areas of north China. Recently, aerosol optical depth (AOD) data have been being used to estimate the PMx concentrations by implementing linear and/or nonlinear regression analysis methods. However, a lot of relevant research based on AOD published so far have demonstrated some limitations in estimating the spatial distribution of PMx concentrations with respect to estimation accuracy and spatial resolution. In this research, the Google Earth Engine (GEE) platform is employed to obtain the band reflectance (BR) data of a large number of Landsat 8 Operational Land Imager (OLI) remote sensing images. Combined with the meteorological, time parameter and the latitude and longitude zone (LLZ) method proposed in this article, a new BR (band reflectance)-PMx (incl. PM10 and PM2.5) model based on a multilayer perceptron neural network is constructed for the estimation of PMx concentrations directly from Landsat 8 OLI remote sensing images. This research used Beijing, China as the test area and the conducted experiments demonstrated that the BR-PMx model achieved satisfactory performances for the PMx-concentration estimations. The coefficient of determination (R2) of the BR-PM2.5 and BR-PM10 models reached 0.795 and 0.773, respectively, and the root mean square error (RMSE) reached 20.09 μg/m3 and 31.27 μg/m3. Meanwhile, the estimation results have been compared with the results calculated by Kriging interpolation at the same time point, and the spatial distribution is consistent. Therefore, it can be concluded that the proposed BR-PMx model provides a new promising method for acquiring accurate PMx concentrations for various cities of China.

Graphical Abstract

1. Introduction

With the rapid development of the economy in China, the processes of industrialization and urbanization have increased the environmental burden, and air pollution has become increasingly serious. Aerosols, which not only have an impact on global climate change but also on the environmental quality of the atmosphere and human health, have become the primary pollutant affecting ambient air quality in most parts of China. Simultaneously, it has been shown that increasing the concentration of aerosol particles is an important reason for the frequent “haze” weather in cities and suburban areas [1,2,3]. Therefore, the estimation of aerosol PMx (including PM10 and PM2.5) has become a popular research topic in recent years [4,5,6,7]. Aerosols have the characteristics of continuous occurrence and large spatial variability in concentration. Due to the uneven distribution of ground monitoring stations, it is difficult to obtain the accurate distributions of aerosol data together with their change trends for an entire city by the methods of spatial interpolation and/or numerical simulation based on the data from monitoring stations [8,9]. Taking advantages of high timeliness, wide coverage and high resolution, the technique of satellite remote sensing makes it possible to monitor the aerosol conditions on a larger spatial scale [10,11,12].
Research on PMx estimation based on satellite remote sensing images began in the 1990s [13]. With the rapid improvement of sensor spectral detection capabilities and the appearance of imaging spectrometers, the satellite data of Moderate Resolution Imaging Spectroradiometer (MODIS) [14], Multi-angle Imaging Spectroradiometer (MISR) [15], Geostationary Meteorological Satellite-5 (GMS-5) [16], Huanjing-1(HJ-1) [17], Himawari-8 [18] and Landsat 8 Operational Land Imager (OLI) have been used for aerosol retrieval studies in the past decades [19].
Meanwhile, a number of previous studies have been conducted to investigate the quantitative relationships between AOD (aerosol optical depth) and PMx concentrations [20,21,22]. The AOD-PMx model is constructed by linear or nonlinear regression, which can be employed to accurately estimate ground-level particulate matter, partly by accounting for relative humidity (RH), wind speed (WS), temperature (TEMP) and planetary boundary layer height (PBLH) [23,24,25]. More recently, Artificial Neural Networks (ANNs) have proven to be an advanced mathematical model to achieve better regression results compared with simple linear/nonlinear regression analysis [26,27].
Although some AOD-PMx models can be used to estimate PMx well in some areas, it is still essential to evaluate the robustness of these methods. Paciorek and Liu [28]. highlighted the limitations of AOD in predicting the spatial distribution of PM2.5; Kumar [29] summarized some of the factors that affect the AOD-PM2.5 association, such as mismatch in spatial-temporal resolution, decomposition of AOD by aerosol types, collocation of AOD and PM2.5 data, and control for spatial–temporal structure in the statistical model. In view of the relative weaknesses between AOD and PM2.5 in the study of Paciorek and Liu [28], some of the findings need further analysis.
This research is dedicated to developing a so-called BR (band reflectance)-PMx model based on the algorithm of a multilayer perceptron neural network for the estimation of the PMx (including PM10 and PM2.5) concentrations, where the various meteorological factors and the time parameter are jointly considered. As a result, the PMx-concentration estimations with high spatial resolution and complete spatial coverage have been realized.

2. Materials and Methods

According to several previous research works concerning the aerosol retrievals from satellite remote sensing images [19,30], the band 1 (0.43–0.45 µm), band 3 (0.53–0.59 µm) and band 7 (2.11–2.29 µm) of Landsat 8 OLI have been considered to be relevant to PMx concentrations. Considering that the band reflectance can be affected by the land surface reflectance, normalized difference vegetation index (NDVI) has been also employed in the proposed BR-Model to realize a more reasonable evaluation of the atmospheric contribution to this reflectance. In this research, the Google Earth Engine (GEE) cloud computing platform has been employed to obtain a large number of consecutive historical satellite image data (Landsat 8 OLI) over the whole area of China from May 2014 to May 2018. Meanwhile, (a) the hourly PMx concentration extracted from the China Environmental Monitoring Center (CEMC) and (b) the meteorological parameters, such as relative humidity (RH), temperature (TEMP), wind direction (WD), wind speed (WS) and atmospheric pressure (PRS) have also been implemented in this research. It should be mentioned that the data of (b) (viz. the meteorological parameters) were initially captured by China Meteorological Administration and had been sent to the National Centers for Environmental Information of National Oceanic and Atmospheric Administration (NOAA) U.S. Department of Commerce for further disseminations. In order to facilitate the description, these meteorological parameters will be termed as ‘NOAA data’ in the following paragraphs of this article. Moreover, a new method has been developed to explore the ‘optimal’ latitude and longitude range of the subset for ANN trainings. This method is termed as an LLZ (latitude and longitude zone) method and can be employed by the proposed BR-PMx model to enhance the estimation precision of the PMx concentrations in different research areas.

2.1. Data Collection and Preprocessing

As illustrated in Table 1, three types of data are employed in this research: (a) Landsat 8 OLI remote sensing images from the Google Earth Engine (GEE) platform, (b) hourly PMx concentrations from the China Environmental Monitoring Station (CEMS) and (c) meteorological parameters of RH, PRS, TEMP, WD and WS from the National Oceanic and Atmospheric Administration (NOAA).

2.1.1. Landsat 8 OLI Data

The Landsat 8 satellite, which carries the Operational Land Imager (OLI) and the Thermal Infrared Sensor (TIRS), was successfully launched by NASA on 11 February 2013. The OLI covers all nine bands of the ETM+ sensor. Due to the generic nature of higher resolution and wider band coverage, the Landsat 8 data have been widely used in Earth resource exploration, natural disaster, agriculture, forestry, animal husbandry management and environmental pollution monitoring.
The Google Earth Engine (GEE) cloud computing platform is a facility for processing remote sensing images in bulk under Google Inc. in Mountain View, CA, USA. As one of the most advanced platforms for the analysis and visualization of geographical data, GEE is able to provide users with a large amount of historical satellite image data for consecutive years (e.g., the MODIS, Landsat, etc.). Meanwhile, GEE also provides algorithms and tools for processing and analyzing petabyte data for information mining [31,32,33]. As one significant subset of Landsat 8 OLI/TIRS data, the “ USGS Landsat 8 Collection 1 Tier 1” has been employed as the primary data for the estimation of PMx concentrations in this study considering that (a) the data in this subset can accurately calibrate the top-of-atmosphere (TOA) reflectance and (b) its temporal resolution is 16 days in China, which is shorter than some other subsets of Landsat 8 OLI/TIRS. Then, a JavaScript program was developed in GEE to obtain, adjust and correct the data from thousands of Landsat 8 OLI images by the method of orthorectification and geographical registration. After that, the selected images are processed using the following steps: (a) remove cloud by the function of mask algorithm (FMask) provided by GEE, (b) establish a circular buffer around each CEMC monitoring station, where the radius of the buffer is 15 meters, viz. the half resolution of Landsat 8 OLI data (30 m), (c) extract the band reflectance of bands 1, 3 and 7 from Landsat 8 OLI images of all the pixels falling inside each established buffer, (d) calculate NDVI from band 4 and band 5 of the Landsat 8 OLI data and (e) calculate the average values of the band reflectance and NDVI in each established buffer. After the data processing from step (a) to (e), the calculated average values together with the imaging time properties were assigned to their corresponding CEMC monitoring stations, which can be regarded as the primary source data for the proposed BR-PMx models. It should be mentioned that the imaging time property will be used as temporal reference in the following processes.

2.1.2. PMx Concentration Data of the China Environmental Monitoring Center (CEMC)

Hourly PMx data can be collected from the China urban air quality real-time release platform (http://106.37.208.233:20035/) maintained by the China Environmental Monitoring Center (CEMC). As one of the public institutions under the Ministry of Ecological Environment, CEMC undertakes the primary tasks of national environmental monitoring. The widespread monitoring stations enable continuous aerosol measurements by using the methods of DUSTTRAK DRX and TEOM [34], which sample ambient air hourly and provide simultaneous measurements of PM2.5 and PM10.
Considering the reliability of data quality [35], the research time interval is set from 1 May 2014 to 1 May 2018. As depicted in Figure 1, there are 1497 CEMC monitoring stations in total distributed within the study area. A time buffer interval of ±0.5 h was used to match the dataset of CEMC ground observations with the Landsat 8 OLI satellite transit phase. Thus, the data of 99,404 PMx observations were compiled.

2.1.3. Other Influence Parameters

PMx concentration has been demonstrated to be affected by pollutant emission and meteorological situations, which exhibit significant time-varying characteristics [36]. Pollution source parameters are not discussed due to the limitations of the data source. In this study, two types of parameters were involved: the meteorological data and time data. According to the imaging time property, the Landsat 8 OLI data is assigned a ‘month number (1, 2, …, 12)’. The ‘month number’ will be regarded as a time parameter in the BR-PMx models, allowing seasonal influences to be taken into account. The meteorological data are primarily archived from the National Center for Environmental Information of NOAA, which involve qualified daily, monthly, seasonal, and yearly measurements of temperature, wind, precipitation, etc. The daily measurement produces data every 3 h starting at 0:00 UTC (viz.8:00 am Beijing time), of which the nearest neighbor time method was used for data selection. The meteorological variables selected in the proposed BR-PMx model are RH, TEMP, WD, WS, and atmospheric pressure (PRS). Over the study area (ref. Figure 1), there are 372 NOAA stations with 949,531 meteorological observations during the period.
In the process of geospatial analysis, it is essential to ensure that all of the relevant elements are at the same or very approximate temporal stamps. The imaging time property of the Landsat 8 OLI data is regarded as the temporal reference for the conducted research. Due to the different time periods of the CEMC and NOAA measurements, the time buffer intervals of ±0.5 h and ±1 h are settled for data from CEMC and NOAA, respectively, to obtain the required parameter datasets. In this way, the data from Landsat 8 OLI, CEMC and NOAA can be harmonized together.

2.2. Methodology

Figure 2 shows the workflow used to estimate the PMx concentrations, which requires four groups of source data, viz. (a) the ‘time parameter’ extracted from Landsat 8 OLI, (b) the preprocessed Band 1, 3, 7 and NDVI from Landsat 8 OLI, (c) meteorological parameters (viz. RH, PRS, TEMP, WD, WS) at NOAA stations, and (d) PMx (PM2.5 and PM10) monitoring values at CEMC stations.
Before inputting to the proposed BR-PMx model, the data of (b) and (c) need to be integrated together by the Near Analysis Algorithm (NAA) (Section 2.2.1, below). The estimation of the PMx concentrations are primarily based on multilayer perceptron (MLP) neural network (Section 2.2.2, below). In order to enhance the estimation accuracy for the PMx concentrations in an entire city, the latitude and longitude zone method (Section 2.2.3, below) has been developed as well as employed by the proposed BR-PMx model, which can help to settle the ‘optimal’ region of the ‘MLP neural network training’ for the specific research area.

2.2.1. Near Analysis Algorithm (NAA)

Due to the different number of monitoring stations and geographic locations, it is complex to directly obtain the meteorological information around the CEMC stations. Therefore, the collocation of various meteorological parameters of the investigated CEMC stations has become a priority condition for PMx estimation.
To integrate the CEMC observations and NOAA datasets into the processes of estimating PMx concentrations seamlessly, the NAA has been employed to compute the distance from each feature in the analytic objects (viz. the ‘Input Elements’ of CEMC) to the nearest features in the neighboring objects (viz. the ‘Proximity Elements’ of NOAA), within the settled search radius (e.g., 100 km). Thus, the nearest dataset of the monitoring stations can be calculated. Then, the necessary meteorological variables around one CEMC station can be acquired from its nearest NOAA station.

2.2.2. Multilayer Perceptron (MLP) Neural Network

Physical models are considered to be the most suitable predictors of small-scale PMx concentrations. However, the complexity of physical models and some other influential factors make it hard to achieve the macroscopic estimation of PMx. Statistical regression and ANN approaches are considered a promising method for estimating PMx concentrations.
A neural network is an important technology for pattern recognition and machine learning. It is the basis of deep learning that simulates the neural network of the human brain to realize artificial intelligence. It has been widely used for diverse remote sensing applications [37,38]. The neural network consists of many interconnected processing units. These units are usually linearly arranged into groups, which can be simulated by electronic circuits or computer programs. Multilayer perceptron (MLP) is a forward-structured ANN, where the input vectors are mapped to output vectors, and the representative models involve a multilayer BPN network [39], RBF network [40], the Hopfield model [41] and so on.
In this paper, the BPN (back-propagation network) is selected for PMx concentration estimation. BPN is a multilayered network for weights’ training with forward nonlinear differentiable functions. The BPN neural network consists of input layer, multiple hidden layers and the output layer (see Figure 2). Each of these layers is connected to all of the cells in the adjacent layer. Simultaneously, there is no connection between the cells of the same layer. Once a pair of learning samples is provided to the network, the activation value of the neuron propagates from the input layer to the output layer, and the neurons in the output layer obtain the input response of the network. After that, the connection weights are modified layer by layer from the output layer through the middle layer and then back to the input layer.
In Figure 3, X i is the input of node i of the input layer, i = 1 N , W i j is the weight between node j of the hidden layer and the node i of the input layer, j = 1 p , θ j is the threshold of the node j of the hidden layer; θ k is the threshold of the node k of the hidden layer; ϕ ( X ) is the excitation function of the hidden layer; W j k is the weight between node k of the hidden layer and node j of the hidden layer, k = 1 q . W k l is the weight between node l of the output layer and node k of the hidden layer, l = 1 M . O l represents the threshold of node l of the output layer, l = 1 M , φ ( X ) is the excitation function of the output layer, and Y ( PM X ) is the output layer node.
In this research, a 4-layer BPN neural network has been developed to estimate the PMx concentration. In the process of determining the input and output factors of the model, (a) the band reflectance of band 1, 3, 7 and the band value of NDVI at the CEMC monitoring stations from the Landsat 8 OLI images are considered as the ‘band parameter’; (b) the NAA method has been utilized to obtain the meteorological data (viz. RH, PRS, TEMP, WD and WS) around the CEMC stations based on the NOAA data; and (c) the dataset of ‘actual’ values of PMx concentration have been measured by CEMC stations. In addition, the latitude and longitude zone (LLZ) method has been developed to get the optimal training region (Section 2.2.3, below). The input variables of the proposed BPN model refer to the ‘band parameter’, the meteorological data and the corresponding time parameter, while the output variables are the measured PMx concentrations at CEMC stations. To improve the computational speed and PMx concentration estimation accuracy, the input and output data must be normalized uniformly to the data range of [−1,1].
For the topologic design of the neural network, the tansig function (tan-sigmoid transfer function, see Figure 4a) is utilized for transmitting the neurons of each layer [42]. The purelin function (linear transfer function, see Figure 4b) is employed as the transfer function of the output layer [43,44]. The number of intermediate nodes of the hidden layers is enlarged from 5 to 20 in a stepwise fashion using the Kolmogorov theorem to establish the nodes of the network structure [45]. Based on the growth method of the network structure, the best number of nodes of the hidden layers is selected as (15, 15) for the continuous training of the network.
In the process of neural network operation, the input data are randomly divided into three subsets. The first subset is for training (60%), the second is for validation (20%), and the third is for testing (20%). The data in the training subset are used to train and adjust the weights on the four-layer neural networks. The validation subset is required to minimize overfitting. The data in the testing subset are implemented to predict the PMx concentration data by the BPN model, for which the optimal connection weights should be determined during the training process. The performance of the established neural network model can be primarily evaluated according to the factors of root mean square error (RMSE) and coefficient of determination (R2) between the predicted and the observed value.

2.2.3. Latitude and Longitude Zone (LLZ) Method

As mentioned earlier, the Latitude and Longitude Zone (LLZ) method can be implemented to compute the ‘optimal’ training region for one specific research area (viz. one city), which can be characterized by six steps:
(a)
the geometric center of the research area is settled as the center point of the latitude and longitude zone (LLZ);
(b)
the initial latitude and longitude width can be calculated by Equation (1):
( L o n w i d t h , L a t w i d t h ) = ξ ( L o n a r e a , L a t a r e a )
where L o n w i d t h and L a t w i d t h respectively represent the longitude and latitude width of the initial LLZ, L o n a r e a and L a t a r e a represent the longitude and latitude widths of the minimum enclosing rectangle (MER) of the research area; and ξ is a coefficient that can be empirically settled between 0.5 and 1.
(c)
the LLZ is progressively growing at 0.1 degrees along both of the longitude and latitude directions;
(d)
the monitoring datasets falling inside the LLZ are selected and inputted to the proposed BR-PMx models for the MLP Neural Network training;
(e)
The eigenvalue R2 will be calculated after the MLP Neural Network training, which can reflect the current accuracy of the PMx concentration estimation;
(f)
the steps (c) to (e) will be continuously operated until the area of LLZ becomes K times larger than the MER of the specific research area, where K can be empirically settled as 2, 3, …N.
(g)
the R2 of all of the LLZs will be compared to each other, and then the LLZ with the largest R2 will be picked out as the ‘optimal’ training region of the specific research area.
As illustrated by Figure 5, the ‘optimal’ training region of LLZ is relevant to distribution of the monitoring stations, which could be smaller, larger or quite similar to the MER of the research area. It is worth mentioning that it is not desired if the LLZ is much smaller than the MER of the research area. Such undesired conditions could be avoided by the empirical coefficient ξ settled in step (b).

3. Results

3.1. PMx Estimations Based on MLP Neraul Network in the Whole Area of Mainland China

Based on the MLP neural network, the PMx concentrations have been estimated in the whole area of mainland China. Several previous studies have demonstrated the temporal patterns of PMx concentrations, i.e., the PMx concentrations may vary with months and/or seasons [46,47]. Hence, the R2 between PMest and PMrea is calculated for different months (from January to December) and seasons (from spring to winter). Then, the scatter plots of PMest and PMrea are drawn from the 70,393 samples for the BR-PM2.5 model and 67,265 samples for the BR-PM10 model by the line regression functions provided by MATLAB 2016, MathWorks Inc, Natick, USA (see Figure 6). With the BR-PM2.5 model, the R2 is 0.3647, and the root mean square error (RMSE) is 25.34 (μg/m3), while with the BR-PM10 model, the R2 is 0.2928 and the RMSE is 36.66 (μg/m3). Obviously, neither result is satisfactory.
To further investigate the temporal aerosol variations, the proposed BR-PMx models are implemented to estimate the concentrations of PM2.5 and PM10 in each month and season from 2014 to 2018. The relevant regression coefficients of the estimation results are illustrated in Table 2. For all monthly and seasonal regression calculations, the R2 of the BR-PM2.5 model is generally slightly higher than that of the BR-PM10 model, while the RMSE of the BR-PM2.5 model is much smaller. For the BR-PM2.5 model, R2 exhibited a seasonal behavior characterized by a maximum during autumn (0.4280), lower values in spring (0.3910) and summer (0.3813), and a minimum in winter (0.3754), which roughly agrees with the findings of Engel-Cox et al. [48]. For the BR-PM10 model, the maximum value of R2 also exists in autumn (0.3407), followed by winter (0.3233), summer (0.2866) and spring (0.2646).

3.2. Caculation of the Optimal LLZ for the Specific Research Area

To improve the accuracy of PMx concentration estimation and reduce the values of RMSE, the LLZ method (described in Section 2.2.2) is integrated into the MLP neural network algorithm in the proposed BR-PMx models. The capital of China, Beijing (116.39 E, 39.92 W) is selected as the test area, and the LLZ center point is set at the geometric center of Beijing, growing at 0.1 degrees along the longitude and latitude direction. Then, the monitoring datasets inside the latitude and longitude zone are continuously tested by the BR-PMx models to explore the optimal bandwidth. Figure 7 and Figure 8 show the variation of the coefficient of determination (R2) between PMest and PMrea at different spatial ranges obtained by the LLZ method. It should be noted that the origin point of the x , y axis is (116.39, 39.92) in both Figure 7 and Figure 8.
According to Figure 7 and Figure 8, it is found that the variation of the R2 between PMest and PMrea is not very drastic in a range with the longitude width within 2° and the latitude width within 1°. Once the latitude/longitude range is enlarged to a certain width, the R2 values decrease dramatically. The optimal values for the coefficient R2 of the BR-PMX model exceed 0.7 with a longitude range of 2° and a latitude range of 1°.
Using the LLZ method, the data located inside the bandwidth of ca. 2° longitude and ca. 1° latitude with Beijing’s geometric center as the center point are inputted to the BR-PMx models. Both the BR-PM10 and BR-PM2.5 models demonstrate significant improvements for the calculated estimates of the PMx concentrations after the combination with the LLZ method. The R2 of PM2.5 and PM10 concentrations increases to 0.795 and 0.773, respectively. The RMSE is reduced by 14.6% (36.65 to 31.27) and 20.7% (25.34 to 20.09) for PM10 and PM2.5, respectively (see Figure 9), which indicates that the LLZ method is very helpful for estimating the PMx concentrations.

3.3. Validation Analysis

As depicted in Section 2.1.1, the remote sensing images imported to the BR-PMx model need to be processed for removing of clouds. For the mainland of China, winter is the season with the most serious air pollution, while the air quality is much better in summer. Then, the data of (a) Landsat 8 OLI remote sensing images at UTC 2:54 (viz. 10:54 am Beijing time), on 30 December 2016 (winter) (see Figure 10, below) and UTC 2:54, on 24 August 2016 (summer) (see Figure 11, below) in Beijing and (b) NOAA meteorological data for the same region and at the approximate time points, are selected and retained. All the other data in December and August of 2016 were removed from the BR-PMx models.
To obtain the continuous distribution graph of the PMx concentrations of the research area, one of the common methods is to operate the spatial interpolation based on the discrete PMx values at various points. Figure 10b,d and Figure 11b,d are such spatial distribution graphs calculated by the Kriging interpolations based on the monitoring PMx values from the 12 CEMC stations located in the area of Beijing, China. Figure 10c,e and Figure 11c,e are the spatial distribution graphs of the PM10 and PM2.5 concentrations estimated by the proposed BR-PMx models in this research. All of these graphs have the spatial resolution of 30 m.
The estimation results of the BR-PMx models and the interpolation results are consistent with respect to the spatial distribution. The higher PMx values are distributed in the south, while the lower values are located in the north, which are not only related to the special terrain of Beijing but also to the industrialization and population density. Furthermore, the minimum (Min), maximum (Max) and the mean values of the PMx concentrations, which are respectively generated by Kriging interpolations and the BR-PMx models, have been illustrated as well as compared in Table 3 and Table 4 (below). Table 3 and Table 4 demonstrate that, under both of the conditions with (a) heavy air pollution at UTC 2:54 on 30 December 2016 and (b) mild air pollution at UTC 2:54 on 24 August 2016, the coefficient of determination (R2) between Kriging interpolation and BR-PM2.5 model reached 0.80817 and 0.82954 respectively, while the R2 of PM10 are respectively 0.77442 and 0.75519: the distribution trend of the PMx concentrations graph generated by Kriging interpolation and BR-PMx models remained consistent regardless of whether the air is heavily polluted or not.
After the validation analysis, it can be concluded that the proposed BR-PMx models are feasible for the PMx concentration estimations with the air pollution in different degrees, and thereby have a certain potential for real-world applications.

4. Discussion

Taking advantage of the wide spatial and temporal coverage, it has been generally proved that the various satellite remote sensing images could be employed for the estimation of the PMx concentrations in large geographic areas. In order to obtain the continuous spatial distribution of PMx concentrations in various cities of China, this research develops new BR-PMx models for the estimation of PMx (including PM10 and PM2.5) concentrations based on the algorithm of a multilayer perceptron neural network combined with LLZ method. In the BR-PMx models, the primary source data were from Landsat 8 OLI and in the meantime, various meteorological factors and the time parameter have been taken into consideration. To the best of our knowledge, the Landsat 8 OLI satellite data have been seldom investigated for the PMx concentration estimations in the literature published so far.
With an explore process with stepwise enlarged areas, the optimal LLZ for Beijing has been identified. Then, the coefficient of determination (R2) of the BR-PM2.5 and BR-PM10 models respectively reached 0.795 and 0.773; synchronously, the RMSE values of the BR-PM2.5 and BR-PM10 models reduced to 31.27 μg/m3 and 20.09 μg/m3, respectively. The estimation accuracy is satisfactory and higher than many established models based on the AOD method [19,20,24,49]. To make the validation analysis, the BR-PMx models have been implemented to obtain the spatial distribution graph of the PMx concentrations in Beijing at two different times: one is in winter with heavily polluted air and the other is in summer with much better air quality. The distribution graphs generated by the BR-PMx models are significantly better than that of Kriging interpolations in terms of resolution and clarity.
Furthermore, as the Landsat 8 OLI remote sensing images have been employed as the primary source data in this research, the distribution graphs of PMx concentrations estimated by the PMx models have a much higher spatial resolution than many other research works, e.g., the estimated PMx concentrations calculated by the AOD method in [19,20,21,24,50], which is based on the MODIS data with the highest resolution of 250 m.
Therefore, the proposed BR-PMx models can be used as one of the possible methods to estimate the general spatial distributions of PMx concentrations in various cities in China that have been monitored by CEMC stations.

5. Conclusions

In this research, we developed a new BR-PMx model to estimate the PMx concentration, which is based on the spectral data of Landsat 8 OLI remote sensing images and the PMx concentration data from the ground monitoring stations. During the process of analyzing the correlations between PMest and PMrea in terms of temporal dimensions with the data from May 2014 to May 2018, satisfactory estimation results cannot be obtained if the whole area of China is set as an estimation region and the regression coefficient R2 between PMest and PMrea for the models BR-PM2.5 and BR-PM10 just reach ca. 0.3647 and 0.2928, respectively. The estimation results represent typical characteristics of seasonal variations. The period with the highest estimation accuracy is in autumn for both the BR-PM2.5 and BR-PM10 models.
To improve the estimation performance of the BR-PMx model, the factor of latitude and longitude zone (LLZ) has been integrated into the model to obtain the optimal training area for the specific research area of Beijing, China. The results demonstrate that the LLZ method performs promisingly as the coefficient of determination (R2) of the BR-PM2.5 and BR-PM10 models reached 0.795 and 0.773, respectively; their RMSE reached 31.27 μg/m3 for the BR-PM2.5 model and 20.09 μg/m3 for the BR-PM10 model.
In general, based on the Landsat 8 OLI remote sensing images and the PMx concentration data, the MLP neural network combined with the LLZ method produced reasonable estimation results of PMx concentration in the test area of Beijing, China. As known, there are some deficiencies in the LLZ method due to the nonuniform distribution of the observation stations. In addition, it is difficult to determine the optimal bandwidth accurately since the optimal bandwidth can vary from one area to another. Due to the fact that the regional characteristics of aerosols can be taken into account by the LLZ method, the LLZ method is very useful to solve the problems of missing or having few observation stations in the research area by acquiring the training data from the surrounding stations. The BR-PMx models after the training in the optimal LLZs can then be utilized to analyze the air pollution characteristics of the investigated research areas. Meanwhile, the high-resolution (30 m) estimations of PMx concentrations with complete spatial coverage were derived in this research. From the perspective of daily coverage, the estimation results are quite close to the observation values. Hence, the estimation results can help to (a) understand the formation process of regional PMx pollution episodes, (b) obtain accurate PMx sources and distributions with high spatial resolution and (c) establish pollution control measures.
In this research, the meteorological factors of RH, TEMP, WD, WS and PRS were considered in the developed BR-PMx models. In our future work, some other parameters (e.g., the planetary boundary layer) and the sensitivity analysis of the employed parameters will be further investigated. Moreover, it should make some sense if the AOD method and the proposed BR-PMx models can be integrated together.

Author Contributions

B.Z., M.Z. and X.Z. conceived and designed the study; B.Z. and J.K. collected and analyzed the data; B.Z. completed the experiments and software; M.Z. and X.Z. initiated the overall research question and M.Z. found funding for this project; B.Z. and M.Z. drafted the paper; D.H. and J.X. critically read and revised the draft. All authors read and approved the final manuscript.

Funding

This research was funded by (a) the National Natural Science Foundation of China, Grant No. 41871315, and (b) the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation program, Grant No.ERC-2016-StG-714087.

Acknowledgments

This research is supported by (a) the Helmholtz Association under the framework of the Young Investigators Group “SiPEO” (VH-NG-1018, www.sipeo.bgu.tum.de), (b) the Bavarian Academy of Sciences and Humanities in the framework of Junges Kolleg, and (c) the Data Center for Resources and Environmental Sciences, Chinese Academy of Sciences (RESDC) (http://www.resdc.cn) as the data provider.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Geng, G.; Zhang, Q.; Martin, R.V.; Donkelaar, A.V.; Huo, H.; Che, H.; Lin, J.; He, K. Estimating long-term PM 2.5 concentrations in China using satellite-based aerosol optical depth and a chemical transport model. Remote Sens. Environ. 2015, 166, 262–270. [Google Scholar] [CrossRef]
  2. King, M.D.; Kaufman, Y.J.; Menzel, W.P.; Tanré, D. Remote sensing of cloud, aerosol, and water vapor properties from the moderate resolution imaging spectrometer (MODIS). IEEE Trans. Geosci. Remote Sens. 1992, 30, 2–27. [Google Scholar] [CrossRef]
  3. Peng, J.; Chen, S.; Lü, H.; Liu, Y.; Wu, J. Spatiotemporal patterns of remotely sensed PM 2.5 concentration in China from 1999 to 2011. Remote Sens. Environ. 1999, 174, 109–121. [Google Scholar] [CrossRef]
  4. Gauderman, W.J.; Avol, E.; Gilliland, F.; Vora, H.; Thomas, D.; Berhane, K.; McConnell, R.; Kuenzli, N.; Lurmann, F.; Rappaport, E.; et al. The effect of air pollution on lung development from 10 to 18 years of age. N. Engl. J. Med. 2004, 351, 1057–1067. [Google Scholar] [CrossRef] [PubMed]
  5. 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. Environ. Health 2013, 12–43. [Google Scholar] [CrossRef]
  6. Pascal, M.; Falq, G.; Wagner, V.; Chatignoux, E.; Corso, M.; Blanchard, M.; Host, S.; Pascal, L.; Larrieu, S. Short-term impacts of particulate matter (PM 10, PM 10–2.5, PM 2.5) on mortality in nine French cities. Atmos. Environ. 2014, 95, 175–184. [Google Scholar] [CrossRef]
  7. Pope, C.A.; Burnett, R.T.; Thun, M.J.; Calle, E.E.; Krewski, D.; Ito, K.; Thurston, G.D. Lung cancer, cardiopulmonary mortality, and long-term exposure to fine particulate air pollution. J. Am. Med. Assoc. 2002, 287, 1132–1141. [Google Scholar] [CrossRef]
  8. You, W.; Zang, Z.; Zhang, L.; Li, Z.; Chen, D.; Zhang, G. Estimating ground-level PM10 concentration in northwestern China using geographically weighted regression based on satellite AOD combined with CALIPSO and MODIS fire count. Remote Sens. Environ. 2015, 168, 276–285. [Google Scholar] [CrossRef]
  9. Christodoulakis, J.; Varotsos, C.A.; Cracknell, A.P.; Kouremadas, G.A. The deterioration of materials as a result of air pollution as derived from satellite and ground based observations. Atmos. Environ. 2018, 185, 91–99. [Google Scholar] [CrossRef] [Green Version]
  10. 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. [Google Scholar] [CrossRef]
  11. Zou, B.; Pu, Q.; Bilal, M.; Weng, Q.; Zhai, L.; Nichol, J.E. High-Resolution Satellite Mapping of Fine Particulates Based on Geographically Weighted Regression. IEEE Geosci. Remote Sens. Lett. 2017, 13, 495–499. [Google Scholar] [CrossRef]
  12. Zou, B.; Zheng, Z.; Wan, N.; Qiu, Y.; Wilson, J.G. An optimized spatial proximity model for fine particulate matter air pollution exposure assessment in areas of sparse monitoring. Int. J. Geogr. Inf. Sci. 2016, 30, 727–747. [Google Scholar] [CrossRef]
  13. Mukai, S.; Sano, I.; Takashima, T.; Yamaguchi, Y. Aerosol retrieval based on photopolarimetry. In Proceedings of the International Geoscience and Remote Sensing Symposium (IGARSS), Pasadena, CA, USA, 8–12 August 1994; Volume 4, pp. 1898–1900. [Google Scholar]
  14. Afonin, S.V. An appraisal of the method of AOD retrieval over land according to MODIS satellite measurements in IR spectral range. Atmos. Ocean. Opt. 2011, 24, 584–586. [Google Scholar] [CrossRef]
  15. Moon, T.; Wang, Y.; Liu, Y.; Yu, B. Evaluation of a MISR-based high-resolution aerosol retrieval method using AERONET DRAGON campaign data. IEEE Trans. Geosci. Remote Sens. 2015, 53, 4328–4339. [Google Scholar] [CrossRef]
  16. Zhang, J.; Si, Z.; Mao, J.; Wang, M. Remote Sensing Aerosol Optical Depth over China with GMS-5 Satellite. Chin. J. Atmos. Sci. 2003, 27, 23–35. [Google Scholar] [CrossRef]
  17. He, J.; Zha, Y.; Zhang, J.; Gao, J.; Li, Y.; Chen, X. Retrieval of aerosol optical thickness from HJ-1 CCD data based on MODIS-derived surface reflectance. Int. J. Remote Sens. 2015, 36, 882–898. [Google Scholar] [CrossRef]
  18. Ge, B.; Yang, L.; Chen, X.; Li, Z.I.; Mei, X.; Liu, L. Study on aerosol optical depth retrieval over land from Himawari-8 data based on dark target method. J. Remote Sens. 2018, 38–50. [Google Scholar] [CrossRef]
  19. Li, R.; Sun, L. An Improved DDV Method to Retrieve Aerosol Optical Depth for Landsat 8 OLI Image. Remote Sens. Inf. 2017, 32, 120–125. [Google Scholar] [CrossRef]
  20. Koelemeijer, R.; 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]
  21. Liu, Y.; Franklin, M.; Kahn, R.; Koutrakis, P. Using aerosol optical thickness to predict ground-level PM 2.5 concentrations in the St. Louis area: A comparison between MISR and MODIS. Remote Sens. Environ. 2007, 107, 33–44. [Google Scholar] [CrossRef]
  22. Nicolantinio, W.D.; Cacciari, A.; Bolzacchini, V.; Ferrero, V.; Volta, M.; Pisoni, E. MODIS aerosol optical properties over North Italy for estimating surface-level PM2.5. In Proceedings of the Envisat Symposium 2007, Montreux, Switzerland, 23–27 April 2007. [Google Scholar]
  23. Gupta, P.; Christopher, S.A. Particulate matter air quality assessment using integrated surface, satellite, and meteorological products: Multiple regression approach. J. Geophys. Res. 2009, D14205. [Google Scholar] [CrossRef]
  24. 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]
  25. 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]
  26. Hong, D.; Yokoya, N.; Xu, J.; Zhu, X. Joint & progressive learning from high-dimensional data for multi-label classification. In Proceedings of the European Conference on Computer Vision (ECCV), Munich, Germany, 8–14 September 2018; pp. 469–484. [Google Scholar]
  27. Xu, J.; Schüssler, O.; Rodriguez, D.G.L.; Romahn, F.; Doicu, A. A Novel Ozone Profile Shape Retrieval Using Full-Physics Inverse Learning Machine (FP-ILM). IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2017, 10, 5442–5457. [Google Scholar] [CrossRef]
  28. 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] [Green Version]
  29. Kumar, N. What Can Affect AOD–PM2.5 Association. Environ. Health Perspect. 2010, 118, A109–A110. [Google Scholar] [CrossRef] [PubMed]
  30. Guo, G.; Ma, L. Urban aerosol optical thickness retrieval from MODIS data. Remote Sens. Technol. Appl. 2005, 20, 343–345. [Google Scholar] [CrossRef]
  31. Dong, J.; Xiao, X.; Menarguez, M.A.; Zhang, G.; Qin, Y.; Thau, D.; Biradar, C.; Moore, B. Mapping paddy rice planting area in northeastern Asia with Landsat 8 images, phenology-based algorithm and Google Earth Engine. Remote Sens. Environ. 2016, 185, 142–154. [Google Scholar] [CrossRef] [Green Version]
  32. Krishnan, S.P.T.; Gonzalez, J.L. Building Your Next Big Thing with Google Cloud Platform; A Press: Berkeley, CA, USA, 2015; pp. 28–49. ISBN 978-1-4842-1004-8. [Google Scholar]
  33. Patel, N.N.; Angiuli, E.; Gamba, P.; Gaughan, A.; Lisini, G.; Stevens, F.R.; Tatem, A.J.; Trianni, G. Multitemporal settlement and population mapping from Landsat using Google Earth Engine. Int. J. Appl. Earth Obs. 2015, 35, 199–208. [Google Scholar] [CrossRef] [Green Version]
  34. Rivas, I.; Mazaheri, M.; Viana, M.; Moreno, T.; Cifford, S.; He, C.; Bischof, O.F.; Martins, V.; Reche, C.; Alastuey, A.; et al. Identification of technical problems affecting performance of DustTrak DRX aerosol monitors. Sci. Total Environ. 2017, 584–585, 849–855. [Google Scholar] [CrossRef] [PubMed]
  35. Ghanem, D.; Zhang, J. Effortless Perfection: Do Chinese Cities Manipulate ‘Blue Skies?’. SSRN Electron. J. 2013, 68, 203–225. [Google Scholar] [CrossRef]
  36. 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]
  37. Zhu, X.X.; Tuia, D.; Mou, L.; Xia, G.S.; Zhang, L.; Xu, F.; Fraundorfer, F. Deep Learning in Remote Sensing: A Comprehensive Review and List of Resources. IEEE Geosci. Remote Sens. Mag. 2017, 5, 8–36. [Google Scholar] [CrossRef] [Green Version]
  38. Hong, D.; Liu, W.; Su, J.; Pan, Z.; Wang, G. A novel hierarchical approach for multispectral palmprint recognition. Neurocomputing 2015, 151, 511–521. [Google Scholar] [CrossRef]
  39. Lee, S.; Choeh, J.Y. Predicting the helpfulness of online reviews using multilayer perceptron neural networks. Expert Syst. Appl. 2014, 41, 3041–3046. [Google Scholar] [CrossRef]
  40. Zhu, Q.; Cai, Y.; Liu, L. A global learning algorithm for a RBF network. Neural Netw. 1999, 12, 527–540. [Google Scholar] [CrossRef]
  41. Farhat, N.H.; Psaltis, D.; Prata, A.; Paek, E. Optical Implementation of The Hopfield Model. Appl. Opt. 1985, 24, 1469. [Google Scholar] [CrossRef] [PubMed]
  42. Ramadhas, A.S.; Jayaraj, S.; Muraleedharan, C.; Padmakumari, K. Artificial neural networks used for the prediction of the cetane number of biodiesel. Renew. Energy 2006, 31, 2524–2533. [Google Scholar] [CrossRef]
  43. Meng, X.; Jia, M.; Wang, T. Neural network prediction of biodiesel kinematic viscosity at 313 K. Fuel 2014, 121, 133–140. [Google Scholar] [CrossRef]
  44. Rashidi, A.M.; Hayati, M.; Rezaei, A. Prediction of the relative texture coefficient of nanocrystalline nickel coatings using artificial neural networks. Solid State Sci. 2011, 13, 1589–1593. [Google Scholar] [CrossRef]
  45. Gardner, M.W.; Dorling, S.R. Artificial neural networks (the multilayer perceptron)—A review of applications in the atmospheric sciences. Atmos. Environ. 1998, 32, 2627–2636. [Google Scholar] [CrossRef]
  46. Liu, J.; Li, J.; Li, W. Temporal Patterns in Fine Particulate Matter Time Series in Beijing. Sci. Rep. 2016, 6, 32221. [Google Scholar] [CrossRef] [PubMed]
  47. Ming, L.; Jin, L.; Li, J.; Fu, P.; Yang, W.; Liu, D.; Zhang, G.; Wang, Z.; Li, X. PM2.5 in the Yangtze River Delta, China: Chemical compositions, seasonal variations, and regional pollution events. Environ. Pollut. 2017, 223, 200–212. [Google Scholar] [CrossRef]
  48. 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]
  49. 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 aerosol optical depth to estimate daily PM2.5 geographical distribution in Mexico City. Environ. Sci. Technol. 2015, 49, 8576–8584. [Google Scholar] [CrossRef]
  50. 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]
Figure 1. Research area and the spatial distribution of CEMC (red points) and NOAA (yellow points) stations used in this study.
Figure 1. Research area and the spatial distribution of CEMC (red points) and NOAA (yellow points) stations used in this study.
Remotesensing 11 00646 g001
Figure 2. The workflow of PMx concentration estimations.
Figure 2. The workflow of PMx concentration estimations.
Remotesensing 11 00646 g002
Figure 3. The diagram of the back-propagation network (BPN).
Figure 3. The diagram of the back-propagation network (BPN).
Remotesensing 11 00646 g003
Figure 4. Transfer functions utilized in the BPN: (a) the tansig function and (b) the purelin function.
Figure 4. Transfer functions utilized in the BPN: (a) the tansig function and (b) the purelin function.
Remotesensing 11 00646 g004
Figure 5. The diagrammatic sketch of the ‘optimal’ LLZ which could be (a) smaller, (b) larger or (c) quite similar to the MER of the specific research area; rectangle with solid line: MER of the research area; rectangle with dashed line: optimal LLZ; blue dots: the monitoring stations.
Figure 5. The diagrammatic sketch of the ‘optimal’ LLZ which could be (a) smaller, (b) larger or (c) quite similar to the MER of the specific research area; rectangle with solid line: MER of the research area; rectangle with dashed line: optimal LLZ; blue dots: the monitoring stations.
Remotesensing 11 00646 g005
Figure 6. Scatterplots between PMest and PMrea of (a) the BR-PM2.5 model and (b) BR-PM10 model in the whole area of mainland China, where the solid line represents the fit line obtained.
Figure 6. Scatterplots between PMest and PMrea of (a) the BR-PM2.5 model and (b) BR-PM10 model in the whole area of mainland China, where the solid line represents the fit line obtained.
Remotesensing 11 00646 g006
Figure 7. The coefficient of determination (R2) between PMest and PMrea of the BR-PM2.5 model at different spatial obtained by the LLZ method.
Figure 7. The coefficient of determination (R2) between PMest and PMrea of the BR-PM2.5 model at different spatial obtained by the LLZ method.
Remotesensing 11 00646 g007
Figure 8. The coefficient of determination (R2) between PMest and PMrea of the BR-PM10 model at different spatial ranges obtained by the LLZ method.
Figure 8. The coefficient of determination (R2) between PMest and PMrea of the BR-PM10 model at different spatial ranges obtained by the LLZ method.
Remotesensing 11 00646 g008
Figure 9. Scatterplots between PMest and PMrea of (a) BR-PM10 model and (b) BR-PM2.5 models combined with the LLZ method, where the solid line is the fitting line obtained by the line regression method provided by MATLAB functions and the dashed line is the reference of the linear function y = x.
Figure 9. Scatterplots between PMest and PMrea of (a) BR-PM10 model and (b) BR-PM2.5 models combined with the LLZ method, where the solid line is the fitting line obtained by the line regression method provided by MATLAB functions and the dashed line is the reference of the linear function y = x.
Remotesensing 11 00646 g009
Figure 10. Spatial distributions of the PM10 and PM2.5 concentrations (μg/m3) over the Beijing area at UTC 2:54 on 30 December 2016: (a) the Landsat 8 OLI image of RGB true color; (b) the PM10 concentration calculated by the method of Kriging interpolation; (c) the results estimated by the BR-PM10 model; (d) the PM2.5 concentration calculated by the method of Kriging interpolation; and (e) the results estimated by the BR-PM2.5 model.
Figure 10. Spatial distributions of the PM10 and PM2.5 concentrations (μg/m3) over the Beijing area at UTC 2:54 on 30 December 2016: (a) the Landsat 8 OLI image of RGB true color; (b) the PM10 concentration calculated by the method of Kriging interpolation; (c) the results estimated by the BR-PM10 model; (d) the PM2.5 concentration calculated by the method of Kriging interpolation; and (e) the results estimated by the BR-PM2.5 model.
Remotesensing 11 00646 g010aRemotesensing 11 00646 g010b
Figure 11. Spatial distributions of the PM10 and PM2.5 concentrations (μg/m3) over the Beijing area at UTC 2:54 on 24 August 2016: (a) the Landsat 8 OLI image of RGB true color; (b) the PM10 concentration calculated by the method of Kriging interpolation; (c) the results estimated by the BR-PM10 model; (d) the PM2.5 concentration calculated by the method of Kriging interpolation; and (e) the results estimated by the BR-PM2.5 model.
Figure 11. Spatial distributions of the PM10 and PM2.5 concentrations (μg/m3) over the Beijing area at UTC 2:54 on 24 August 2016: (a) the Landsat 8 OLI image of RGB true color; (b) the PM10 concentration calculated by the method of Kriging interpolation; (c) the results estimated by the BR-PM10 model; (d) the PM2.5 concentration calculated by the method of Kriging interpolation; and (e) the results estimated by the BR-PM2.5 model.
Remotesensing 11 00646 g011
Table 1. Details of all datasets used in this study.
Table 1. Details of all datasets used in this study.
CategoryVariablesUnitTemporal ResolutionSource
PMx concentrationPM2.5
PM10
μg/m31 hChina Environmental Monitoring Center (CEMC)
Satellite image data of Landsat 8 Operational Land Imager (OLI)Band1
Band3
Band7
NDVI
Band Reflectance16 dayGoogle Earth Engine (GEE)
Meteorological parametersRH%3 hNational Oceanic and Atmospheric Administration (NOAA)
PRSPa3 h
TEMPK3 h
WD°3 h
WSm/s3 h
Table 2. Regression coefficients of BR-PMx models using the samples in each month/season from 2014 to 2018.
Table 2. Regression coefficients of BR-PMx models using the samples in each month/season from 2014 to 2018.
PeriodBR-PM2.5BR-PM10
R2RMSENR2RMSEN
March0.314128.3568310.318655.836597
April0.292623.2751950.220839.995078
May0.358823.5745800.260642.454438
June0.381720.5951720.306837.384896
July0.340521.0654040.261325.775128
August0.347621.256990.268931.775377
September0.311215.3553740.292821.235111
October0.381428.0162900.342339.296011
November0.348230.3857670.282952.675505
December0.391440.5768950.360347.196626
January0.387740.663960.380553.166069
February0.345330.9168190.302849.056428
ALL Spring0.381320.68166060.264637.7716113
ALL Summer0.391017.28162750.286628.3415401
ALL Autumn0.428025.4174310.340739.7716627
ALL Winter0.375433.78201100.323350.0119141
Table 3. Comparison of the PMx-concentration distributions generated by Kriging interpolation and the BR-PMx models at UTC 2:54 on 30 December 2016 (in winter).
Table 3. Comparison of the PMx-concentration distributions generated by Kriging interpolation and the BR-PMx models at UTC 2:54 on 30 December 2016 (in winter).
CategoryMin (μg/m3)Max (μg/m3)Mean (μg/m3)R2 between (a) and (b)
(a) PM2.5 Kriging interpolation46.43316.41127.970.80817
(b) Estimation by BR-PM2.5 Model24.92336.42145.88
(a) PM10 Kriging interpolation105.67271.91174.110.77442
(b) Estimation by BR-PM10 Model71.45323.52169.46
Table 4. Comparison of the PMx-concentration distributions generated by Kriging interpolation and the BR-PMx models at UTC 2:54 on 24 August 2016 (in summer).
Table 4. Comparison of the PMx-concentration distributions generated by Kriging interpolation and the BR-PMx models at UTC 2:54 on 24 August 2016 (in summer).
CategoryMin (μg/m3)Max (μg/m3)Mean (μg/m3)R2 between (a) and (b)
(a) PM2.5 Kriging interpolation33.7665.1947.540.82954
(b) Estimation by BR-PM2.5 Model14.2687.3138.65
(a) PM10 Kriging interpolation51.66101.8769.220.75519
(b) Estimation by BR-PM10 Model23.12125.3853.74

Share and Cite

MDPI and ACS Style

Zhang, B.; Zhang, M.; Kang, J.; Hong, D.; Xu, J.; Zhu, X. Estimation of PMx Concentrations from Landsat 8 OLI Images Based on a Multilayer Perceptron Neural Network. Remote Sens. 2019, 11, 646. https://doi.org/10.3390/rs11060646

AMA Style

Zhang B, Zhang M, Kang J, Hong D, Xu J, Zhu X. Estimation of PMx Concentrations from Landsat 8 OLI Images Based on a Multilayer Perceptron Neural Network. Remote Sensing. 2019; 11(6):646. https://doi.org/10.3390/rs11060646

Chicago/Turabian Style

Zhang, Bo, Meng Zhang, Jian Kang, Danfeng Hong, Jian Xu, and Xiaoxiang Zhu. 2019. "Estimation of PMx Concentrations from Landsat 8 OLI Images Based on a Multilayer Perceptron Neural Network" Remote Sensing 11, no. 6: 646. https://doi.org/10.3390/rs11060646

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