Next Article in Journal
Post-Processing Ensemble Precipitation Forecasts and Their Applications in Summer Streamflow Prediction over a Mountain River Basin
Next Article in Special Issue
Application of Trigonometric Polynomial Fitting Method in Simulating the Spatial Distribution of PM2.5 Concentration in South-Central China
Previous Article in Journal
A Low-Cost Wireless Sensor Network for Barn Climate and Emission Monitoring—Intermediate Results
Previous Article in Special Issue
Spatiotemporal Variation and Driving Factors for NO2 in Mid-Eastern China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Soil-Derived Dust PM10 and PM2.5 Fractions in Southern Xinjiang, China, Using an Artificial Neural Network Model

1
School of Geographic and Environmental Sciences, Tianjin Normal University, Tianjin 300387, China
2
State Key Laboratory of Environmental Criteria and Risk Assessment, Chinese Research Academy of Environmental Science, Beijing 100012, China
3
Tianjin Research Institute for Water Transport Engineering, M.O.T., Tianjin 300456, China
4
Department of Planning, Industry and Environment, New South Wales Government, Parramatta, NSW 7057, Australia
5
College of Computer Science, Nankai University, Tianjin 300071, China
6
Tianjin Changhai Environmental Monitoring Service Corporation, Tianjin 300000, China
*
Authors to whom correspondence should be addressed.
Atmosphere 2023, 14(11), 1644; https://doi.org/10.3390/atmos14111644
Submission received: 20 July 2023 / Revised: 22 September 2023 / Accepted: 28 September 2023 / Published: 31 October 2023
(This article belongs to the Special Issue Air Quality and the Implementation of Sustainable Development Goals)

Abstract

:
Soil-derived dust emissions have been widely associated with health and environmental problems and should therefore be accurately and reliably estimated and assessed. Of these emissions, the inhalable PM10 and PM2.5 are difficult to estimate. Consequently, to calculate PM10 and PM2.5 emissions from soil erosion, an approach based on an artificial neural network (ANN) model which provides a multilayered, fully connected framework that relates input parameters and outcomes was proposed in this study. Owing to the difficulty in obtaining the actual emissions of soil-derived PM10 and PM2.5 over a broad area, the PM10 and PM2.5 simulated results of the ANN model were compared with the published results simulated by the widely used wind erosion prediction system (WEPS) model. The PM10 and PM2.5 emission results, based on the WEPS, agreed well with the field data, with R2 values of 0.93 and 0.97, respectively, indicating the potential for using the WEPS results as a reference for training the ANN model. The calculated r, RMSE and MAE for the results simulated by the WEPS and ANN were 0.78, 3.37 and 2.31 for PM10 and 0.79, 1.40 and 0.91 for PM2.5, respectively, throughout Southern Xinjiang. The uncertainty of the soil-derived PM10 and PM2.5 emissions at a 95% CI was (−66–106%) and (−75–108%), respectively, in 2016. The results indicated that by using parameters that affect soil erodibility, including the soil pH, soil cation exchange capacity, soil organic content, soil calcium carbonate, wind speed, precipitation and elevation as input factors, the ANN model could simulate soil-derived particle emissions in Southern Xinjiang. The results showed that when the study domain was reduced from the entire Southern Xinjiang region to its five administrative divisions, the performance of the ANN improved, producing average correlation coefficients of 0.88 and 0.87, respectively, for PM10 and PM2.5. The performances of the ANN differed by study period, with the best result obtained during the sand period (March to May) followed by the nonheating (June to October) and heating periods (November to February). Wind speed, precipitation and soil calcium carbonate were the predominant input factors affecting particle emissions from wind erosion sources. The results of this study can be used as a reference for the wind erosion prevention and soil conservation plans in Southern Xinjiang.

1. Introduction

Wind erosion has been recognized as a major environmental problem in arid and semiarid regions [1,2], with soil erosion from wind threatening air quality by emitting dust into the atmosphere. Globally, annual soil-derived, wind-blown atmospheric dust is estimated to be as high as 3000 million tons [3]. Soil-derived dust that travels long distances can be a major source of atmospheric particles [4,5]. The PM10 and PM2.5 fractions associated with dust emissions are of particular concern because they can be easily inhaled and thus cause adverse health effects such as respiratory, cardiovascular and dermatological disorders, and even death [6,7,8]. It was found that dust-related PM10 emissions from the Saharan Desert were significantly correlated with hospital and daily mortality in 13 Southern European cities [9]. Researchers measured long-term (2012–2020) PM10 concentrations in a city in the Sistan Basin, Eastern Iran, which is one of the dustiest arid environments globally, and found that the quality of air for 25% of the days was hazardous to people’s health [10].
Several studies have shown that soil-derived particle emissions from wind erosion sources are influenced by various meteorological factors [11,12,13], with wind and precipitation being the dominant drivers [13,14]. Dust particles, such as PM10 and PM2.5, are usually produced by a mechanism known as the saltation process, which requires the wind regime to reach a critical wind velocity (threshold wind speed) sufficient to dislodge particles from the soil surface [15,16]. Higher soil moisture from increased precipitation allows capillary forces to develop between soil grains, reinforcing soil cohesion and inhibiting wind erosion [17]. Soil-derived particulate emissions are also associated with soil surface properties [18,19]. For example, calcium carbonate (CaCO3) and organic content (OC) can affect the amount of dust generated from the soil surface by changing the soil cohesion [20,21,22].
Various methods have been used to study the wind erosion and emission properties of particles, namely field investigations, wind tunnel experiments and model simulations (Table S1). Compared with field measurements and wind tunnel experiments, wind erosion modeling, based on field measurements and geospatial data, is a more efficient and effective method to estimate dust emissions from soil wind erosion [23]. The wind erosion prediction system (WEPS), developed by the US Department of Agriculture, is one of the most widely used models for estimating soil erosion and particle production from soil surfaces and has been extensively applied throughout the United States [24,25,26,27,28] and in countries worldwide, such as Canada [29], Argentina [30] and China [31,32,33]. In numerical studies, good agreement (i.e., R2 ranging from 0.87 to 0.98) was obtained between the measured and the WEPS-based simulated erosion results [34,35]. However, this model requires detailed information related to the PM10 emission process, including meteorology, hydrology, land management and soil surface erodibility data. The collection and input of such large datasets are time and cost consuming and may restrict the use of comprehensive simulation models [31,36].
With the development of big data and artificial intelligence technology, different machine learning models such as LSTM-INFO, ELM-SAMOA, RVM-IMRFO and RVFL-QANA were used in the field of soil erosion [37,38,39]. As a commonly used model, artificial neural networks (ANNs) are universal and can learn complex nonlinear relationships between important input and output factors by approximating a large class of functions with a high degree of accuracy [40,41]. It has been successfully used in the field to explain soil erosion phenomena and does not require the restrictive assumptions of the WEPS model [42,43]. In a complex system of soil erosion, a full understanding of the erosion regime and transport process is challenging, although the difficulties in simulating such complex phenomena could be significantly reduced through the application of ANN models [2]. However, as shown in Table S2, studies related to ANN-based soil erosion have mainly focused on simulating the total soil loss [44,45,46] with few studies using ANN models for soil-derived dust emissions related to PM10 and PM2.5 fractions. This is probably due to the difficulty in obtaining, through field measurements, data on actual emissions of this particulate matter at a regional scale. Because of this challenge, an approach was proposed for estimating soil-derived particle emissions by using an ANN model and comparing its results with the WEPS results obtained from our prior work [33]; this was verified by the PM emissions data collected from field measurements undertaken during this study. The use of ANNs in estimating soil-derived PM10 and PM2.5 emissions will benefit governments in the implementation of appropriate control measures to improve regional air quality and protect public health in arid and semiarid areas.
The aim of this study was to select, optimize and apply an ANN method to estimate soil-derived particles (PM10 and PM2.5) from wind erosion sources. The ANN-estimated emissions from the wind-erosion sources were compared with the results obtained from the WEPS model. A further aim was to specify the importance of factors affecting soil erosion via a sensitivity analysis technique (ANN and Monte Carlo combined model). The significance of our study demonstrated the successful use of an ANN combined with the WEPS model to simulate soil-derived PM10 and PM2.5 when the selected input parameters are available; this avoids the substantial effort required to obtain particle emissions data at a regional scale through field measurements.

2. Materials and Methods

2.1. Study Area

The study area, Southern Xinjiang (73.4–88.2° E, 34.2–42.3° N), is approximately 1.06 million km2 and comprises five administrative divisions (Kashi, Kezhou, Hetian, Bazhou and Akesu; Figure 1). A semiarid or desert climate prevails in Southern Xinjiang, which has substantial seasonal differences in temperature, with hot summers and cold winters. The annual average temperature is 12.4 °C, with mean temperatures in summer and winter of 25.1 °C and −3.3 °C, respectively. The annual average precipitation is 59.1 mm. The average wind speed at a 10 m height is 1.8 m/s; the highest wind speed (2.15 m/s) is in spring, followed by summer (2.14 m/s), autumn (1.44 m/s) and winter (1.29 m/s). The Taklamakan Desert, which is unique and serves as the main wind erosion source of particulate matter, is situated in the middle of Southern Xinjiang and is the second largest drifting desert globally. In 2016, daily concentrations of PM10 (PM2.5) averaged 435 μg/m3 (158 μg/m3), 211 μg/m3 (77 μg/m3), 285 μg/m3 (106 μg/m3), 103 μg/m3 (42 μg/m3) and 243 μg/m3 (92 μg/m3) for Kashi, Kezhou, Hetian, Bazhou and Akesu, respectively (China National Environmental Monitoring Center, http://www.cnemc.cn/, accessed on 3 October 2023). These average daily concentrations of PM10 and PM2.5 in Southern Xinjiang exceeded the National Ambient Air Quality Standard of China (GB3095-2012), (https://www.transportpolicy.net/standard/china-air-quality-standards/, accessed on 3 October 2023), (daily average of 70 μg/m3 for PM10 and 35 μg/m3 for PM2.5) by 3.65 and 2.71 times, respectively.

2.2. Data Collection and Soil Properties Analysis

A total of 264 representative soil samples were collected from the study area. As shown in Figure 1, the samples were evenly distributed around the Taklamakan Desert because obtaining samples from the desert hinterland was difficult. The samples were collected from the top 10 cm of the soil surface by cutting rings, which were then transported to the laboratory for analysis of their characteristics, including their pH, CaCO3, cation exchange capacity (CEC) and OC. These factors were selected because they have a strong influence on wind erosion [47,48].
Two meteorological parameters including wind speed (WS) at a 10 m height and precipitation (Pre) were collected from the American National Center for Atmospheric Research (https://ncar.ucar.edu/, accessed on 3 October 2023), which were then downscaled by the Weather Research and Forecasting model resampled to a resolution of 10 km. An elevation (Ele) map of the study area was generated by using a digital elevation model from the Geospatial Data Cloud (http://www.gscloud.cn/, accessed on 3 October 2023).

2.3. Artificial Neural Network Model

The structure of the ANN model used in this study is shown in Figure 2. The total study area was gridded into 17,576 data cells with a resolution of 10 km × 10 km. The cells were divided randomly into two subsets: 70% of data cells for training and the remaining 30% for testing. First, the training datasets were used to train the model to obtain the optimal weight and network structure. Next, the determined optimal ANN model structure and input factors in the testing datasets were used to simulate the soil-derived particles in areas without erosion (the output in the testing datasets). To evaluate the performance of the ANN in simulating soil-derived particles from wind erosion in areas with unknown erosion amounts, the ANN’s simulated values were compared with the corresponding reference values obtained by the WEPS published in our previous study [33].
In addition, the study year was divided into three periods: the sand period (March–May), nonheating period (June–October) and heating period (November–February) [49,50]. Data from the last month in each period were used as the testing dataset, with data from the other months being used as the training dataset. The values predicted by the ANN in May, October and February were further compared with their corresponding reference values, obtained through the WEPS, in terms of evaluating the model performance of the ANN in predicting soil-derived particles from wind erosion.
As shown in Figure 2, the network structure was designed in the form of 7-5-1: one input layer with seven neurons, one hidden layer with five neurons and one output layer that represented the soil-derived dust emissions related to PM10 or PM2.5. The seven neurons were the soil pH, soil CaCO3, soil CEC, soil OC, WS, Pre and elevation. Previous studies have reported that an ANN with one hidden layer can estimate any nonlinear problem with high accuracy [51,52]; therefore, one hidden layer was set for our model. The number of neurons in the hidden layer was determined based on the two-thirds rule [53]; i.e., two-thirds of the number of neurons in the input layer (7 × 2/3 ≈ 5). The normalized input factors were applied to two commonly used activation functions, namely the logarithmic sigmoid function and hyperbolic tangent function, to introduce nonlinearity into the output. Detailed information is shown in the section headed “Principal of artificial neural network model” in the Supplementary Materials. The ANN model analysis was conducted by using the “neuralnet” package in R software RoxygenNote 6.1.0 (https://cran.r-project.org/web/packages/neuralnet/neuralnet.pdf, accessed on 3 October 2023).

2.4. Field Experiment

Because the results obtained by the WEPS were used as a reference for training the ANN model, these WEPS results had to be evaluated by using field measurement data from the selected study area. To measure the wind erosion flux of the particles between October 2016 and March 2017, field experiments were conducted by using a Sensit (Redlands, CA, USA) H14-LIN wind erosion mass flux sensor at a national field observation site (80°43′45″ E, 37°00′57″ N) in the Hetian administrative division. The erosion sensor was installed at a height of 2.5 m above the ground, and daily average particle concentrations with sizes ranging from 50 to 70 μm were measured by the instrument. The emissions of PM10 and PM2.5 were calculated based on the following equations:
E P M 10 = α E P M 50 70
E P M 2.5 = β E P M 50 70
where EPM10, EPM2.5 and EPM50–70 represent the amount of soil-derived particles with sizes less than or equal to 10 μm; 2.5 μm; and the amount of eroded particles collected by the sensor, with particle sizes ranging from 50 to 70 μm, respectively. α and β are constant factors, which were set as 0.11 and 0.03, respectively.
The constant factors were calculated based on the information in Table S3, which was obtained from a study which analyzed the particle size distribution from sand dust storms in the Taklamakan Desert [54]. The monthly average values of the PM emissions were calculated and compared with the WEPS-simulated monthly values in the cell where the observation site was located.

2.5. Modeling Validation

The predictive ability of the ANN model for estimating soil-derived particles was examined based on several statistical indexes, including the determination coefficient (R2), Pearson correlation coefficient (r), root mean square error (RMSE) and mean absolute error (MAE). The equations of these indexes are shown in the section “Model validation” in the Supplementary Materials. The closer the RMSE and MAE values are to zero, the more accurate the simulations are. When the absolute value of r is in the range of 0.7 to 1, a strong correlation is obtained.

2.6. Uncertainty Analysis and Sensitivity Analysis

The uncertainty analysis of the amount of soil-derived PM10 and PM2.5, obtained by the ANN, was conducted by using the Monte Carlo method [55,56]. The distributions and values of the parameters of these input factors are listed in Table 1. A total of 100,000 samples, without replacement for each input factor, were resampled from the fitted probability distribution. Seven blocks of 100,000 input vectors were used to generate 100,000 values of PM10 (or PM2.5) emissions (t/km2) by using the optimal structure of the ANN. The uncertainty propagation was calculated by using the following equation:
U n c e r t a i n t y = ( V 2.5 % M e a n M e a n × 100 % ~ V 97.5 % M e a n M e a n × 100 % )
where V2.5% and V97.5% are the 2.5th and 97.5th percentile value, respectively, of the generated PM amount. The narrower the 95% confidence interval (CI) obtained by the above equation, the greater the accuracy of the prediction.
The sensitivity analysis was performed to probe the response of the soil-derived particle emissions to independent changes in the soil parameters and meteorological data [57]. Detailed information on the sensitivity analysis is shown in the “Sensitivity analysis” section in the Supplementary Materials. In addition, a Spearman analysis was used to verify the effect of each factor on soil-derived particle emissions.

3. Results

3.1. Verifying PM Emissions of WEPS Based on Field Experiments

During the study, the monthly average values of the PM emissions per unit area were calculated. As shown in Figure 3, the PM10 and PM2.5 emission results based on the WEPS agreed well with the field data, with R2 values of 0.93 and 0.97, respectively. The absolute values of the RMSE were 0.34 and 0.18 for PM10 and PM2.5, respectively. Based on this comparison, the emission results obtained by the WEPS were deemed reliable and acceptable. Thus, using the simulation data obtained by the WEPS as a reference for training the ANN model is possible.

3.2. PM Emissions from Wind Erosion Sources Based on ANN

After network training and optimization, it was found that the model performance with the logarithmic sigmoid function could provide better results than that with the hyperbolic tangent function for both PM10 and PM2.5 (Table 2). Thus, the optimized network structure, which is 7-5-1 with a logarithmic sigmoid function as the activation function, was applied to different administrative divisions and study periods. Test data with unknown erosion amounts were selected randomly by the ANN (30% of the total data) to test the model performance of the ANN in simulating the dust emissions for the study area. Figure 4 shows the violin chart, which compares the results obtained by the ANN by using the optimized network structure and WEPS. It was found that the distribution of the soil-derived dust emissions simulated by the WEPS is compatible with the distribution obtained by the ANN regardless of particle size. As shown in Table 3, the values of R2 ranged from 0.53 to 0.66 for PM10 and 0.57 to 0.66 for PM2.5 throughout Southern Xinjiang, indicating that the model explains 53–66% (57–66%) of the variation in the soil-derived PM10 (PM2.5) around its mean. The values of R were approximately equivalent to 0.8 regardless of the particle size (see Figure S1 in the Supplementary Materials). It was also found that R2 increased when the study domain was reduced. In Kezhou, the maximum R2 reached 0.92 and 0.93 for PM10 and PM2.5, respectively, during the study period. The value of the correlation coefficient ranged from 0.73 to 0.97, indicating a strong positive linear relationship between the results simulated by the ANN and WEPS. In summary, the results indicated that the performance of the ANN model in estimating PM emissions for the reduced study area were more reliable than those for the entire area of Southern Xinjiang.
Moreover, Table 4 shows the performance of the ANN model by using the optimal network architecture to predict soil-derived dust emissions from wind erosion sources in the last month for a period. The results showed that the predictivity of PM emissions by the ANN in the sand and nonheating periods was better than that in the heating period, because a higher R2 and r and lower RMSE and MAE were obtained for the sand and nonheating periods than the values obtained for the heating period.

3.3. Uncertainty Analysis

The uncertainty analysis of the simulated PM emissions from wind erosion sources in Southern Xinjiang was conducted by estimating the 95% CI of the simulated results. As shown in Table 5, the uncertainty of the soil-derived PM10 emissions at a 95% CI was (−66–106%) in 2016. The uncertainty of PM10 during different periods was (−70–68%), (−71–136%) and (−72–115%) for the sand, heating and nonheating periods, respectively. The uncertainties for PM10 were lower than the uncertainties for PM2.5, indicating that the ANN model was more accurate in simulating the emission of PM10 than that of PM2.5. Figure 5 shows the fitted probability distribution function for the PM10 emission inventory, which indicates that the PM10 emissions from wind erosion sources followed a log normal distribution. All the PM10 values were within the 95% CI regardless of the study period, confirming the robustness of the ANN model [57].

3.4. Sensitivity Analysis

Because of the high uncertainty of PM2.5, a sensitivity analysis was only conducted on PM10. The obtained PM10 emission range and the results of the Spearman analysis are presented in Table 1. The value of ΔPM10 was the highest (42.03 t/km2) for Pre, followed by WS (14.86 t/km2) and CaCO3 (11.13 t/km2), indicating that the Pre, WS and CaCO3 of the soil were the most important factors for soil-derived PM10 emissions in Southern Xinjiang. The contributions of the pH, CEC and OC of the soil were relatively low, with the contribution of elevation being the lowest. The relationship between PM10 emissions and the influencing factors was analyzed by estimating the Spearman correlation coefficients between them. The negative value of R2 (R2 = −0.20) showed that with increasing precipitation, PM10 emissions decreased accordingly. Simulated PM emissions increased with an increase in the WS (R2 = 0.36). The growth of the soil CaCO3 and OC resulted in a significant decrease in PM10 emissions.

4. Discussion

A selected ANN model was used to simulate PM10 and PM2.5 from a wind erosion source by using relevant factors, including soil properties and meteorological parameters as input variables. The results indicated that an ANN can be used as an alternative tool to simulate soil-derived particle emissions provoked by wind erosion in areas with similar meteorological conditions and soil properties. The determined optimal network structure was used to simulate the emissions of PM10 and PM2.5 in an area with unknown emission amounts. The high correlation coefficients and low errors in five administrative divisions in Southern Xinjiang indicated that the model was able to simulate PM emissions for places without erosion data. The effectiveness of machine learning models in terms of solving the soil erosion problem was also reported by a study which evaluated dust emissions in the Jazmurian Basin in Southeastern Iran by using six algorithms [58]. All the models yielded a high performance with R > 0.9 and RMSE < 20%. It is also found that for regions such as the entire area of Southern Xinjiang, the performance of the ANN requires further refinement. This phenomenon can be explained by similar meteorological conditions and similar soil properties in a relatively small study area. The contribution of each input parameter to soil-derived dust emissions from wind erosion sources differs by administrative division [59,60]. An ANN was also used to predict the soil-derived particles for the last month of each study period. The ANN model performed better in the sand and nonheating periods compared with that of the heating period. This is probably because the soil samples were collected during the sand and nonheating periods, so the properties of the soil used in the modeling process represented the soil conditions in the sand and nonheating periods.
The Monte Carlo technique was used to examine the uncertainty of the ANN model in simulating PM emissions from soil erosion. The uncertainty analysis also confirmed that the ANN model could properly simulate PM10 emissions from soil in Southern Xinjiang because all the PM10 values fell within the 95% confidence interval. It is found that the ANN model was more accurate in simulating the emission of PM10 than that of PM2.5, probably because the PM2.5 emission inventory was obtained by multiplying the PM10 emissions by factors obtained from the wind tunnel experiments. This process may introduce additional uncertainties to the ANN model. To the best of our knowledge, limited studies were conducted on estimating soil-derived particles (PM10 and PM2.5) by using an ANN. Most studies reported that an ANN can be applied to the modeling of soil erosion rates, which measure the amount of soil mass lost from a unit area. For example, Kouchami-Sardoo et al. studied the performance of an ANN on predicting the soil erosion rate in Kerman province, Southeast Iran, by using factors such as the bulk density, soil organic matter and soil moisture [46]; they found a high correlation (R2 = 0.94, MSE = 0.04) between the erosion rates measured in the wind tunnel and the values simulated by the ANN. Kim et al. studied the soil erosion rate at a farm in Lincoln, USA, by using an ANN model [52]. By comparing the field measurement values, R2 and RMSE were reported as 0.62 and 0.093, respectively. Egbueri et al. evaluated the performance of the ANN for forecasting soil erodibility potentials in Southern Anambra, and they found that the R2 for an ANN model ranged from 0.971 to 0.998 and the sum of residual errors ranged from 0.000 to 0.024 [61]. Thus, given the fact that ANNs can decrease the estimated time and costs, the use of this model in simulating soil erosion is recommended [62].
Although the accuracy of ANN-based simulation results is high, the ANN’s inability to interpret the physical influences of the input parameters on soil erosion is one major disadvantage of ANN-based studies. The combined use of an ANN and Monte Carlo simulation was thus conducted to analyze the sensitivity of each variable to PM emissions, allowing us to assess the importance of each component [46]. Statistical analyses showed that the Pre, WS and CaCO3 of the soil were the three major factors affecting soil erosion. Similarly, studies reported that rainfall was the most important factor in the soil-erosion process [3,58]. The study year had the highest recorded rainfall since 1961 at 247.6 mm. The increased soil moisture due to rainfall creates substantial interparticle forces and increases aggregation between soil particles, which inhibits soil erosion and dust emission [3]. It was found that the influence of moisture on PM10 emissions decreases logarithmically [63]. Wind speed is the most direct power source and a leading factor in soil erosion [64]. Numerous studies have found that PM10 emissions from soil increased significantly with an increase in WS, regardless of soil type [65,66,67]. Coupled with the special topography of the Tarim Basin in Southern Xinjiang, considerable wind shear could be generated because of the Kunlun Mountain block along the southern edge of the Xinjiang region, producing advantageous conditions for high soil-derived dust emissions [33]. A high WS contributes to soil erosion and increases the emission of particles, which can explain why the emissions of PM10 and PM2.5 were the highest during the sand period (throughout Southern Xinjiang, the sand period had the highest average WS of 2.37 m/s). The calcium carbonate content and organic matter of soil are typical cementing materials for aggregation in soils, which limit the emission of particles by wind erosion [68]. Soil aggregation produces a high demand for water, which can fill the intra-aggregate pores and bond the dust particles, inhibiting PM emissions [69]. It was found that calcium carbonate in soil could increase the content of aggregates, which decreases the potential for soil erosion by wind [70]. It was also reported that PM10 emissions were low because of high OC in soils [63].
The limitation of this study is the use of simulation data obtained by the WEPS for training the ANN model. In the absence of validated observational field measurement data, the data produced by globally available WEPS model simulations were used as the data for comparing with the ANN model. The error of the results simulated by the WEPS was not considered. Although field data are the most representative and best data to use to validate the performance of an ANN model, dust emission data with a high spatial resolution were seldom obtained through field experiments in Southern Xinjiang. To mitigate the limitation, a field measurement was carried out at a single sampling site to verify the accuracy of the results simulated by the WEPS. Further field experiments at a regional level, which require a relatively high spatial resolution, are needed to measure actual soil-derived dust emissions related to PM10 and PM2.5. Thus, the efficiency of an ANN in simulating PM emissions from wind erosion sources can be verified directly by measuring actual field data.

5. Conclusions

Our study evaluated the performance of an ANN model in simulating soil-derived dust emissions related to PM10 and PM2.5 fractions in Southern Xinjiang, which includes the second largest drifting desert worldwide. By comparing its results with well-established modeling data obtained by the WEPS model, it was concluded that the ANN model provided acceptable results for highly nonlinear problems such as a soil erodibility analysis, which is complex and has many influencing factors. The calculated r, RMSE and MAE for the results simulated by the WEPS and ANN were 0.78, 3.37 and 2.31 for PM10, respectively. The corresponding values were 0.79, 1.40 and 0.91 for PM2.5, respectively. The results also indicated that when the study domain was reduced from the entire Southern Xinjiang region to its five administrative divisions, the performance of the ANN improved, producing average correlation coefficients for the comparison of 0.88 and 0.87, respectively, for PM10 and PM2.5. This phenomenon is probably due to the strength of each input parameter on PM10 emissions being different in different administrative divisions. The best performance of the ANN model in simulating PM emissions was achieved during the sand period; this was followed by the nonheating and heating periods. The Monte Carlo uncertainty analysis indicated the robustness of the ANN model because all the PM10 emission values fell within the 95% CI regardless of the study period or particle size. Using an ANN to estimate PM10 and PM2.5 emissions due to soil erosion has several advantages; for example, the model requires less input parameters to estimate soil-derived particles as it is flexible and can be updated in terms of different structures to improve the estimation results. A sensitivity analysis using an ANN and Monte Carlo combined model revealed that the WS, Pre and CaCO3 in the soil were the most important factors influencing soil-derived particle emissions. The limitation of this study was that we only compared the soil-derived dust emission results obtained by the WEPS with data collected at one available field measurement site. For future related investigations, it is recommended to increase the number of locations where measurement data are collected for use in verifying the results obtained by an ANN. The method provided in this study can be used as a tool for the prevention of wind erosion and for soil conservation plans.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/atmos14111644/s1, Figure S1: Performance evaluation of the ANN model using a Taylor diagram involving RMSE and correlation coefficient for southern Xinjiang; Table S1: Summary of published studies related to soil-derived dust emission related to PM10 and PM2.5 fractions in recent five years; Table S2: Summary of published ANN studies related to soil erosion; Table S3: Percentage of different sized particles during the sand-storm on 23 April 2014 in Taklimakan Desert [71,72,73,74,75,76,77,78,79,80,81,82,83,84,85].

Author Contributions

Conceptualization, M.A.; methodology, H.Z. (Hong Zhao), H.W., H.Z. (Hui Zhang) and Y.S.; software, H.Z. (Hong Zhao) and J.M.; resources, N.H., Z.M. and L.C.; data curation, J.Y., J.Z. and Y.L.; writing—original draft preparation, S.G.; writing—review and editing, S.G.; supervision, Z.B., W.Y. and L.C.; project administration, Z.B.; funding acquisition, S.G. All authors have read and agreed to the published version of the manuscript.

Funding

The authors thank the National Key Research and Development Program (grant no. 2016YFC0201700) and the National Natural Science Foundation of China (grant no. 41907194) for funding this research.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to privacy restrictions.

Conflicts of Interest

Author Hui Wang was employed by the company Tianjin Shanghai Environmental Monitoring Service Corporation. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Copeland, S.M.; Bradford, J.B.; Duniway, M.C.; Schuster, R.M. Potential impacts of overlapping land-use and climate in a sensitive dryland: A case study of the Colorado Plateau, USA. Ecosphere 2017, 8, e01823. [Google Scholar] [CrossRef]
  2. Raei, B.; Ahmadi, A.; Neyshaburi, M.R.; Ghorbani, M.A.; Asadzadeh, F. Comparative evaluation of the whale optimization algorithm and backpropagation for training neural networks to model soil wind erodibility. Arab. J. Geosci. 2021, 14, 29. [Google Scholar] [CrossRef]
  3. Katra, I. Soil erosion by wind and dust emission in semi-arid soils due to agricultural activities. Agronomy 2020, 10, 89. [Google Scholar] [CrossRef]
  4. Prospero, J.M. Long-range transport of mineral dust in the global atmosphere: Impact of African dust on the environment of southeastern United States. Proc. Natl. Acad. Sci. USA 1999, 96, 3396–3403. [Google Scholar] [CrossRef]
  5. Díaz-Nigenda, E.; Tatarko, J.; Jazcilevich, A.D.; García, A.R.; Caetano, E.; Ruíz-Suárez, L.G. A modeling study of aeolian erosion enhanced by surface wind confluences over Mexico city. Aeolian Res. 2010, 2, 143–157. [Google Scholar] [CrossRef]
  6. Sharratt, B.S.; Lauer, D. Particulate matter concentration and air quality affected by windblown dust in the Columbia Plateau. J. Environ. Qual. 2006, 35, 2011–2016. [Google Scholar] [CrossRef]
  7. Goudie, A.S. Desert dust and human health disorders. Environ. Int. 2014, 63, 101–113. [Google Scholar] [CrossRef]
  8. Stafoggia, M.; Zauli-Sajani, S.; Pey, J.; Samoli, E.; Alessandrini, E.; Basagana, X.; Cernigliaro, A.; Chiusolo, M.; Demaria, M.; Diaz, J.; et al. Desert dust outbreaks in Southern Europe: Contribution to daily OM10 concentrations and short-term associations with mortality and hospital admissions. Environ. Health Perspect. 2016, 124, 413–419. [Google Scholar] [CrossRef]
  9. Jiménez, E.; Linares, C.; Martínez, D.; Díaz, J. Role of Saharan dust in the relationship between particulate matter and short-term daily mortality among the elderly in Madrid (Spain). Sci. Total Environ. 2010, 408, 5729–5736. [Google Scholar] [CrossRef]
  10. Behrooz, R.D.; Mohammadpour, K.; Broomandi, P.; Kosmopoulos, P.G.; Gholami, H.; Kaskaoutis, D.G. Long-term (2012–2020) PM10 concentrations and increasing trends in the Sistan Basin: The role of Levar wind and synoptic meteorology. Atmos. Pollut. Res. 2022, 13, 101460. [Google Scholar] [CrossRef]
  11. Chepil, W.S. Dynamics of wind erosion: 1. Nature of movement of soil by wind. Soil Sci. 1945, 60, 305–320. [Google Scholar] [CrossRef]
  12. Shao, Y. (Ed.) Physics and Modelling for Wind Erosion, Atmospheric and Oceanographic Sciences Library; Springer: Dordrecht, The Netherlands, 2008. [Google Scholar]
  13. Zhang, G.; Azorin-Molina, C.; Shi, P.; Lin, D.; Guijarro, J.; Kong, F.; Chen, D. Impact of near-surface wind speed variability on wind erosion in the eastern agro-pastoral transitional ozone of Northern China, 1982–2016. Agric. For. Meteorol. 2019, 271, 102–115. [Google Scholar] [CrossRef]
  14. Tai, A.; Ma, P.; Chan, Y.; Chow, M.; Ridley, D.; Kok, J. Impacts of climate and land cover variability and trends on springtime East Asian dust emission over 1982–2010: A modeling study. Atmos. Environ. 2021, 254, 118348. [Google Scholar] [CrossRef]
  15. Kok, J.; Parteli, E.; Michaels, T.; Karam, D. The physics of wind-blown sand and dust. Rep. Prog. Phys. 2012, 75, 106901. [Google Scholar] [CrossRef] [PubMed]
  16. Chen, S.; Huang, J.; Qian, Y.; Zhao, C.; Kang, L.; Yang, B.; Wang, Y.; Liu, Y.; Yuan, T.; Wang, T.; et al. An overview of mineral dust modeling over East Asia. J. Meteorol. Res. 2017, 31, 633–653. [Google Scholar] [CrossRef]
  17. Bergametti, G.; Rajot, J.L.; Pierre, C.; Bouet, C.; Marticorena, B. How long does precipitation inhibit wind erosion in the Sahel? Geophys. Res. Lett. 2016, 43, 6643–6649. [Google Scholar] [CrossRef]
  18. Boroughani, M.; Pourhashemi, S.; Gholami, H.; Kaskaoutis, D.G. Predicting of dust storm source by combining remote sensing, statistic-based predictive models and game theory in the Sistan watershed, southwestern Asia. J. Arid. Land 2021, 13, 1103–1121. [Google Scholar] [CrossRef]
  19. He, Q.; Yang, X.; Mamtimin, A.; Tang, S. Impact factors of soil wind erosion in the center of Taklimakan Desert. J. Arid. Land 2011, 3, 9–14. [Google Scholar]
  20. Hevia, G.E.; Mendez, M.J.; Buschiazzo, D.E. Tillage affects soil aggregation parameters linked with wind erosion. Geoderma 2007, 140, 90–96. [Google Scholar] [CrossRef]
  21. Ben-Hur, M.; Yolcu, G.; Uysal, H.; Lado, M.; Paz, A. Soil structure changes: Aggregate size and soil texture effects on hydraulic conductivity under different saline and sodic conditions. Aust. J. Soil Res. 2009, 47, 688–696. [Google Scholar] [CrossRef]
  22. Virto, I.; Gartzia-Bengoetxea, N.; Fernández-Ugalde, O. Role of organic matter and carbonates in soil aggregation estimated using laser diffractometry. Pedosphere 2011, 21, 566–572. [Google Scholar] [CrossRef]
  23. Tatarko, J.; Donk, S.J.; Ascough, J.C., II; Walker, D.G. Application of the WEPS and SWEEP models to non-agricultural disturbed lands. Heliyon 2016, 2, e00215. [Google Scholar] [CrossRef] [PubMed]
  24. Van Donk, S.J.; Skidmore, E.L. Measurement and simulation of wind erosion, roughness degradation and residue decomposition on an agricultural field. Earth Surf. Process. Landf. 2003, 28, 1243–1258. [Google Scholar] [CrossRef]
  25. Hagen, L.J. Evaluation of the wind erosion prediction system (WEPS) erosion sub-model on cropland fields. Environ. Model. Softw. 2004, 19, 171–176. [Google Scholar] [CrossRef]
  26. Feng, G.; Sharratt, B.S. Validation of WEPS for soil and PM10 loss from agricultural fields on the Columbia Plateau of the United States. Earth Surf. Process. Landf. 2007, 32, 743–753. [Google Scholar] [CrossRef]
  27. Chung, S.H.; Herron-Thorpe, F.L.; Lamb, B.K.; VanReken, T.M.; Vaughan, J.K.; Gao, J.; Wagner, L.E.; Fox, F. Application of the wind erosion prediction system in the AIRPACT regional air quality modeling framework. Trans. ASABE 2013, 56, 625–641. [Google Scholar] [CrossRef]
  28. Gao, J.; Wagner, L.E.; Fox, F.; Chung, S.J.; Vaughan, J.K.; Lamb, B.K. Spatial application of WEPS for estimating wind erosion in the Pacific Northwest. Trans. ASABE 2013, 6, 613–624. [Google Scholar]
  29. Coen, G.M.; Tatarko, J.; Martin, T.C.; Cannon, K.R.; Goddard, T.W.; Sweetland, N.J. A method for using WEPS to map wind erosion risk of Alberta soils. Environ. Model. Softw. 2004, 19, 185–189. [Google Scholar] [CrossRef]
  30. Buschiazzo, D.E.; Zobeck, T.M. Validation of WEQ, RWEQ and WEPS wind erosion for different arable land management systems in the Argentinean Pampas. Earth Surf. Process. Landf. 2008, 33, 1839–1850. [Google Scholar] [CrossRef]
  31. Chen, L.; Zhao, H.; Han, B.; Bai, Z. Combined use of WEPS and Models-3/CMAQ for simulating wind erosion source emission and its environmental impact. Sci. Total Environ. 2014, 466–467, 762–769. [Google Scholar] [CrossRef]
  32. Chen, L.; Zhao, H.; Wang, W.; Bai, Z.; Wang, Z.; Sun, F.; Hou, L.; Liu, G.; Shi, M.; Miao, Y. Effect of windblown dust from local and regional sources on the air quality of the central district in Jinan, China. Atmos. Res. 2017, 185, 44–52. [Google Scholar] [CrossRef]
  33. Liu, Y.; Teng, Y.; Liang, S.; Li, X.; Zhao, J.; Shan, M.; Chen, L.; Yu, H.; Mao, J.; Zhang, H.; et al. Establishment of PM10 and PM2.5 emission inventories from wind erosion source and simulation of its environmental impact based on WEPS-Models3 in southern Xinjiang, China. Atmos. Environ. 2021, 248, 118222. [Google Scholar] [CrossRef]
  34. Funk, R.; Skidmore, E.L.; Hagen, L.J. Comparison of wind erosion measurements in Germany with simulated soil losses by WEPS. Environ. Model. Softw. 2004, 19, 177–183. [Google Scholar] [CrossRef]
  35. Liu, B.; Qu, J.; Niu, Q.; Han, Q. Comparison of measured wind tunnel and SWEEP simulated soil losses. Geomorphology 2014, 207, 23–29. [Google Scholar] [CrossRef]
  36. Bhattacharya, B.; Solomatine, D.P. Application of artificial neural network in stage-discharge relationship. In Proceedings of the 4th International Conference on Hydro-Informatics, Iowa City, IA, USA, 23–27 July 2000; pp. 1–7. [Google Scholar]
  37. Tao, H.; Jawad, A.H.; Shather, A.H.; Al-Khafaji, Z.; Rashid, T.A.; Ali, M.; Al-Ansari, N.; Marhoon, H.A.; Shahid, S.; Yaseen, A.M. Machine learing algorithms for high-resolution prediction of spatiotemporal distribution of air pollution from meteorological and soil parameters. Environ. Int. 2023, 175, 107931. [Google Scholar] [CrossRef]
  38. Goyal, P.; Kumar, S.; Sharda, R. A review of the artificial intelligence (AI) based techniques for estimating reference evapotranspiration: Current trends and future perspectives. Comput. Electron. Agric. 2023, 209, 107836. [Google Scholar] [CrossRef]
  39. Lu, Y.; Li, T.; Hu, H.; Zeng, X. Short-term prediction of reference crop evapotranspiration based on machine learning with different decomposition methods in arid areas of China. Agric. Water Manag. 2023, 279, 108175. [Google Scholar] [CrossRef]
  40. Samarasinghe, S. Neural Networks for Applied Sciences and Engineering: From Fundamentals to Complex Pattern Recognition; CRC Press: Boca Raton, FL, USA, 2007. [Google Scholar]
  41. Shirani, H. Artificial Neural Network with an Application in Agricultural and Natural Resource Sciences, 1st ed.; Vali-E-Asr University of Rafsanjan: Rafsanjan, Iran, 2018. [Google Scholar]
  42. Merdun, H.; Cinar, O.; Meral, R.; Apan, M. Comparision of artificial neural network and regression pedotransfer functions for prediction of soil water retention and saturated hydraulic conductivity. Soil Tillage Res. 2006, 90, 108–116. [Google Scholar] [CrossRef]
  43. Beucher, A.; Moller, A.B.; Greve, M.H. Artificial neural networks and decision tree classification for predicting soil drainage classes in Denmark. Geoderma 2017, 352, 351–359. [Google Scholar] [CrossRef]
  44. De la Rosa, D.; Moreno, J.A.; Mayol, F.; Bonsón, T. Assessment of soil erosion vulnerability in Western Europe and potential impact on crop productivity due to loss of soil depth using the ImpelERO model. Africult. Ecosyst. Environ. 2000, 81, 179–190. [Google Scholar] [CrossRef]
  45. Licznar, P.; Nearing, M. Artificial neural networks of soil erosion and runoff prediction at the plot scale. CATENA 2003, 51, 89–114. [Google Scholar] [CrossRef]
  46. Albaradeyia, I.; Hani, A.; Shahrour, I. WEPP and ANN models for simulating soil loss and runoff in a semi-arid Mediterranean region. Environ. Monit. Assess. 2011, 180, 537–556. [Google Scholar] [CrossRef] [PubMed]
  47. Zou, X.; Zhang, C.; Cheng, H.; Kang, L.; Wu, X.; Chang, C.; Wang, Z.; Zhang, F.; Li, J.; Liu, C.; et al. Classification and representation of factors affecting soil wind erosion in a model. Adv. Earth Sci. 2014, 29, 875–889. (In Chinese) [Google Scholar]
  48. Li, Y.; Song, Y.; Kaskaoutis, D.G.; Zan, J.; Orozbaev, R.; Tan, L.; Chen, X. Aeolian dust dynamics in the Fergana Valley, Central Asia, since ~30 ka inferred from loess deposits. Geosci. Front. 2021, 12, 101180. [Google Scholar] [CrossRef]
  49. Kong, S.; Ji, Y.; Lu, B.; Bai, Z.; Chen, L.; Han, B.; Li, Z. Chemical compositions and sources of atmospheric PM10 in heating, non-heating and sand periods at a coal-based city in northeastern China. J. Environ. Monit. 2012, 14, 852. [Google Scholar] [CrossRef]
  50. Yang, J.; Zhao, T.; Cheng, X.; Ren, Z.; Meng, L.; He, Q.; Tan, C.; Zhu, Y.; Zhu, C.; Wu, S. Temporal and spatial variations of sandstorm and the related meteorological influences over northern China from 2000 to 2019. Acta Sci. Circumstantiae 2021, 41, 2966–2975. (In Chinese) [Google Scholar]
  51. Schalkoff, R.J. Artificial Neural Networks; McGraw-Hill: New York, NY, USA, 1997; pp. 146–188. [Google Scholar]
  52. Kim, M.; Gilley, J. Artificial neural network estimation of soil erosion and nutrient concentrations in runoff from land application areas. Comput. Electron. Agric. 2008, 64, 268–275. [Google Scholar] [CrossRef]
  53. Cortina-Januchs, M.G.; Dominguez, J.Q.; Corona, A.V.; Andina, D. Development of a model for forecasting of PM10 concentrations in Salamanca, Mexico. Atmos. Pollut. Res. 2015, 6, 626–634. [Google Scholar] [CrossRef]
  54. Wang, M.; Ming, H.; Huo, W.; Xu, H.; Li, J.; Li, X. Detecting sand-dust storms using a wind-profiling radar. J. Arid. Land 2017, 9, 753–762. [Google Scholar] [CrossRef]
  55. Guan, Y.; Chen, G.; Cheng, Z.; Yan, B.; Hou, L. Air pollutant emissions from straw open burning: A case study in Tianjin. Atmos. Environ. 2017, 171, 155–164. [Google Scholar] [CrossRef]
  56. Hu, L.; Chen, L.; Mao, J.; Sun, Y.; Gao, S.; Yang, W.; Xiao, J.; Zhang, H. Uncertainty analysis of emission inventory of PM10 from open source blowing dust from wind erosion of bare soil in Tianjin. China Environ. Sci. 2020, 40, 41–51. (In Chinese) [Google Scholar]
  57. Gao, M.; Yin, L.; Ning, J. Artificial neural network model for ozone concentration estimation and Monte Carlo analysis. Atmos. Environ. 2018, 184, 129–139. [Google Scholar] [CrossRef]
  58. Gholami, H.; Mohamadifar, A.; Sorooshian, A.; Jansen, J.D. Machine-learning algorithms for predicting land susceptibility to dust emissions: The case of the Jazmurian Basin, Iran. Atmos. Pollut. Res. 2020, 11, 1303–1315. [Google Scholar] [CrossRef]
  59. Gholami, H.; Mohammadifar, A.; Golzari, S.; Kaskaoutis, D.G.; Collins, A.L. Using the Boruta algorithm and deep learning models for mapping land susceptibility to atmospheric dust emissions in Iran. Aeolian Res. 2021, 50, 100682. [Google Scholar] [CrossRef]
  60. Gholami, H.; Mohammadifar, A.; Malakooti, H.; Esmaeilpour, Y.; Golzari, S.; Mohammadi, F.; Li, Y.; Song, Y.; Kaskaoutis, D.G.; Fitzsimmons, K.E.; et al. Integrated modelling for mapping spatial sources of dust in central Asia—An important dust source in the global atmospheric system. Atmos. Pollut. Res. 2021, 12, 101173. [Google Scholar] [CrossRef]
  61. Egbueri, J.C.; Igwe, O.; Omeka, M.E.; Agbasi, J.C. Development of MLR and variedly optimized ANN models for forecasting the detachability and liquefaction potential index of erodible soils. Geosyst. Geoenviron. 2023, 2, 100104. [Google Scholar] [CrossRef]
  62. Gholami, V.; Khaleghi, M.R.; Sebghati, M.R. A method of groundwater quality assessment based on fuzzy network-CANFIS and geographic information system (GIS). Appl. Water Sci. 2016, 7, 3633–3647. [Google Scholar] [CrossRef]
  63. Aimar, S.B.; Mendez, M.J.; Funk, R.; Buschiazzo, D.E. Soil properties related to potential particulate matter emissions (PM10) of sandy soils. Aeolian Res. 2012, 3, 437–443. [Google Scholar] [CrossRef]
  64. Ge, J.M.; Huang, J.P.; Xu, C.P.; Qi, Y.L.; Liu, H.Y. Characteristics of Taklimankan dust emission and distribution: A satellite and reanalysis field perspective. J. Geophys. Res. Atmos. 2014, 119, 11772–11783. [Google Scholar] [CrossRef]
  65. He, W.; Gao, W.; Tuo, D.; Zhao, P. Study on some factors influencing soil erosion by wind tunnel experiment in north ecotone between agriculture and pasture. J. Soil Water Conserv. 2004, 18, 1–8. (In Chinese) [Google Scholar]
  66. Liu, M.; Wang, J.; Yan, P.; Liu, L.; Li, X. Influence of ridge tillage on soil wind erosion on rainfed cropland: A wind tunnel simulation. J. Soil Water Conserv. 2005, 19, 54–57. (In Chinese) [Google Scholar]
  67. Gao, S.; Wang, Y.; Shan, M.; Teng, Y.; Hong, N.; Sun, Y.; Mao, J.; Ma, Z.; Xiao, J.; Azzi, M.; et al. Wind-tunnel and modelled PM10 emissions and dust concentrations from agriculture soils in Tianjin, northern China. Aeolian Res. 2020, 42, 100562. [Google Scholar] [CrossRef]
  68. Mendez, M.J.; de Oro, L.D.; Panebianco, J.E.; Colazzo, J.C.; Buschiazzo, D.E. Organic carbon and nitrogen in soils of semiarid Argentina. J. Soil Water Conserv. 2006, 61, 230–235. [Google Scholar]
  69. Funk, R.; Reuter, H.E.; Hoffmann, C.; Engel, W.; Ottl, D. Effect of moisture on fine dust emission from tillage operations on agricultural soils. Earth Surf. Process. Landf. 2008, 33, 1851–1863. [Google Scholar] [CrossRef]
  70. Chepil, W.S. Influence of moisture on erodibility of soil by wind. Soil Sci. Soc. Am. Proc. 1956, 20, 288–292. [Google Scholar] [CrossRef]
  71. Avecilla, F.; Panebianco, J.E.; Mendez, M.J.; Buschiazzo, D.E. PM10 emission efficiency for agricultural soils: Comparing a wind tunnel, a dust generator, and the open-air plot. Aeolian Res. 2018, 32, 116–123. [Google Scholar] [CrossRef]
  72. De Oro, L.A.; Avecilla, F.; Panebianco, J.E.; Buschiazzo, D.E. PM10 emission from feedlots in soils with different texture: Cattle trampling effect. Aeolian Res. 2021, 53, 100742. [Google Scholar] [CrossRef]
  73. Funk, R.; Papke, N.; Hor, B. Wind tunnel tests to estimate PM10 and PM2.5-emissions from complex substrates of open-cast strip mines in Germany. Aeolian Res. 2019, 39, 23–32. [Google Scholar] [CrossRef]
  74. Gao, S.; Zhao, H.; Bai, Z.; Han, B.; Xu, J.; Zhao, R.; Zhang, N.; Chen, L.; Lei, X.; Shi, W.; et al. Combined use of principal component analysis and artificial neural network approach to improve estimates of PM2.5 personal exposure: A case study on older adults. Sci. Total Environ. 2020, 726, 138533. [Google Scholar] [CrossRef]
  75. Gholami, V.; Sahour, H.; Amri, M. Soil erosion modeling using erosion pins and artificial neural networks. Catena 2021, 196, 104902. [Google Scholar] [CrossRef]
  76. Gholami, V.; Booij, M.J.; Tehrani, E.N.; Hadian, M.A. Spatial soil erosion estimation using an artificial neural network (ANN) and field plot data. Catena 2018, 163, 210–218. [Google Scholar] [CrossRef]
  77. Kouchami-Sardoo, I.; Shirani, H.; Esfandiarpoiur-Boroujeni, I.; Besalatpour, A.A. Prediction of soil wind erodibiltiy using a hybrid Genetic algorithm-artificial neural network method. Catena 2020, 187, 104315. [Google Scholar] [CrossRef]
  78. Lin, X.; Niu, J.; Yu, X.; Berndtsson, R.; Wu, S.; Xie, S. Maize residue effects on PM2.5, PM10, and dust emission from agricultural land. Soil Tillage Res. 2020, 205, 104738. [Google Scholar] [CrossRef]
  79. Lturri, L.A.; Funk, R.; Leue, M.; Sommer, M.; Buschiazzo, D.E. Wind sorting affects differently the organo-mineral composition of saltating and particulate materials in contrasting texture agricultural soils. Aeolian Res. 2017, 28, 39–49. [Google Scholar]
  80. Padoan, E.; Maffia, J.; Balsari, P.; Ajmone-Marsan, F.; Dinuccio, E. Soil PM10 emission potential under specific mechanical stress and particles characteristics. Sci. Total Environ. 2021, 779, 146468. [Google Scholar] [CrossRef] [PubMed]
  81. Pi, H.; Sharratt, B. Evaluation of the RWEQ and SWEEP in simulating soil and PM10 loss from a portable wind tunnel. Soil Tillage Res. 2017, 170, 94–103. [Google Scholar] [CrossRef]
  82. Tatarko, J.; Kucharski, M.; Li, H.; Li, H. PM2.5 and PM10 emissions by breakage during saltation of agricultural soils. Soil Tillage Res. 2020, 208, 104902. [Google Scholar] [CrossRef]
  83. Tian, M.; Gao, J.; Zhang, L.; Zhang, H.; Feng, C.; Jia, X. Effects of dust emissions from wind erosion of soil on ambient air quality. Atmos. Pollut. Res. 2021, 12, 101108. [Google Scholar] [CrossRef]
  84. Van Pelt, R.S.; Baddock, M.C.; Zobeck, T.M.; D’Odorico, P.; Ravi, S.; Bhattachan, A. Total vertical sediment flux and PM10 emissions from disturbed Chihuahuan Deseart surfaces. Geoderma 2017, 293, 19–25. [Google Scholar] [CrossRef]
  85. Zhang, J.; Zhang, C.; Chang, C.; Wang, R.; Liu, G. Comparison of wind erosion based on measurements and SWEEP simulation: A case study in Kangbao County, Hebei province, China. Soil Tillage Res. 2017, 165, 169–180. [Google Scholar] [CrossRef]
Figure 1. Distribution of sampling sites (yellow dots) and land use within the administrative divisions of Southern Xinjiang.
Figure 1. Distribution of sampling sites (yellow dots) and land use within the administrative divisions of Southern Xinjiang.
Atmosphere 14 01644 g001
Figure 2. Schematic diagram of the methodology showing stages used for soil-derived particle emission estimates based on ANN and Monte Carlo analysis.
Figure 2. Schematic diagram of the methodology showing stages used for soil-derived particle emission estimates based on ANN and Monte Carlo analysis.
Atmosphere 14 01644 g002
Figure 3. Correlation between measured PM emissions per unit area by field measurement and simulated PM emissions per unit area by WEPS in Hetian (the blue line represents 1:1 line).
Figure 3. Correlation between measured PM emissions per unit area by field measurement and simulated PM emissions per unit area by WEPS in Hetian (the blue line represents 1:1 line).
Atmosphere 14 01644 g003
Figure 4. Violin chart comparing the simulated soil-derived dust emissions related to PM10 and PM2.5 fractions obtained from ANN and WEPS in Southern Xinjiang during different periods.
Figure 4. Violin chart comparing the simulated soil-derived dust emissions related to PM10 and PM2.5 fractions obtained from ANN and WEPS in Southern Xinjiang during different periods.
Atmosphere 14 01644 g004
Figure 5. Distribution of soil-derived PM10 emissions from wind erosion sources by using Monte Carlo simulation.
Figure 5. Distribution of soil-derived PM10 emissions from wind erosion sources by using Monte Carlo simulation.
Atmosphere 14 01644 g005
Table 1. Distributions for input factors, Spearman correlation coefficients between input factors and PM10 emissions and PM10 ranges using the optimum ANN model.
Table 1. Distributions for input factors, Spearman correlation coefficients between input factors and PM10 emissions and PM10 ranges using the optimum ANN model.
Input FactorsDistributionValues of ParametersCorrelation with PM10 Emissions per Unit Area (R2)ΔPM10 (t/km2)
pHNormalMean = 7.82; Sd = 0.27−0.04 **8.08
CaCO3NormalMean = 0.048; Sd = 0.038−0.23 **11.13
CECLog normalGeometric Mean = 16.67; Geometric Sd = 1.15−0.09 **6.15
OCLog normalGeometric Mean = 0.0099; Geometric Sd = 1.16−0.04 **5.25
WSWeibullShape = 6.29; Scale = 2.210.36 **14.86
PreExponentialRate = 0.0016−0.20 **42.03
EleGeneralized extreme valueLocation = 9.14; Scale = 3.04; Shape = −0.040.04 **1.06
**: p < 0.01.
Table 2. Performance of ANN model using 7-5-1 network structure with different activation functions to predict soil-derived particles from wind erosion in Southern Xinjiang.
Table 2. Performance of ANN model using 7-5-1 network structure with different activation functions to predict soil-derived particles from wind erosion in Southern Xinjiang.
DistrictActive FunctionStructurePM10PM2.5
R2rRMSEMAER2rRMSEMAE
KashiLogarithmic sigmoid7-5-10.750.870.170.090.710.840.110.05
Tangent sigmoid7-5-10.650.800.240.120.630.790.110.06
KezhouLogarithmic sigmoid7-5-10.920.962.881.730.930.970.510.32
Tangent sigmoid7-5-10.870.933.922.740.940.970.430.24
HetianLogarithmic sigmoid7-5-10.800. 902.891.980.800.891.210.83
Tangent sigmoid7-5-10.830.912.781.890.760.871.350.91
BazhouLogarithmic sigmoid7-5-10.740.861.581.090.640. 800.490.25
Tangent sigmoid7-5-10.720.851.621.110.560.700.620.29
AkesuLogarithmic sigmoid7-5-10.710.843.002.060.720.851.310.91
Tangent sigmoid7-5-10.660.813.282.040.760.871.180.79
Table 3. Performance of ANN model using optimal network architecture (7-5-1 network structure with logarithmic sigmoid function) to predict soil-derived particles from wind erosion in area with unknown emissions in Southern Xinjiang during different periods.
Table 3. Performance of ANN model using optimal network architecture (7-5-1 network structure with logarithmic sigmoid function) to predict soil-derived particles from wind erosion in area with unknown emissions in Southern Xinjiang during different periods.
DistrictPeriodPM10PM2.5
R2rRMSEMAER2rRMSEMAE
KashiSand0.680.830.170.100.740.860.070.05
Heating0.690.830.090.060.680.820.070.04
Nonheating0.740.860.150.110.740.860.100.06
Whole year0.750.870.170.090.710.840.110.05
KezhouSand0.790.892.331.180. 900.950.260.16
Heating0.770.880.900.530.820.910.120.07
Nonheating0.820.911.950.990.880.940.250.14
Whole year0.920.962.881.730.930.970.510.32
HetianSand0.810.901.491.030.790.890.630.43
Heating0.550.740.740.520.570.760.260.20
Nonheating0.780.881.451.000.760.870.620.40
Whole year0.800.902.891.980.800.891.210.83
BazhouSand0.620.790.990.530.650.810.210.12
Heating0.740.860.610.430.730.860.140.11
Nonheating0.700.840.590.410.670.820.150.09
Whole year0.740.861.581.090.640.800.490.25
AkesuSand0.640.801.871.180.660.810.820.57
Heating0.710.840.690.470.670.820.330.23
Nonheating0.670.820.900.610.720.850.380.25
Whole year0.710.843.002.060.720.851.310.91
Southern XinjiangSand0.530.731.861.320.620.790.760.49
Heating0.590.770.770.530.570.750.270.18
Nonheating0.660.811.290.820.660.810.520.30
Whole year0.610.783.372.310.620.791.400.91
Table 4. Performance of ANN model by using optimal network architecture (7-5-1 network structure with logarithmic sigmoid function) to predict soil-derived particles from wind erosion in the last month during each period in Southern Xinjiang.
Table 4. Performance of ANN model by using optimal network architecture (7-5-1 network structure with logarithmic sigmoid function) to predict soil-derived particles from wind erosion in the last month during each period in Southern Xinjiang.
DistrictPeriodPM10PM2.5
R2rRMSEMAER2rRMSEMAE
KashiSand0.570.761.671.660.640.800.800.79
Heating0.530.730.060.040.550.740.030.02
Nonheating0.550.740.080.030.58−0.760.110.09
KezhouSand0.950.970.920.810.910.950.160.14
Heating0.680.820.060.040.640.800.050.03
Nonheating0.900.951.180.730.880.940.200.13
HetianSand0.760.870.720.590.740.860.300.25
Heating0.610.780.170.100.520.720.080.06
Nonheating0.680.830.650.250.590.760.380.25
BazhouSand0.790.890.750.700.690.720.070.05
Heating0.680.820.280.200.520.830.190.17
Nonheating0.720.840.520.300.470.680.130.08
AkesuSand0.710.840.900.790.740.860.410.36
Heating0.560.750.430.310.570.750.200.14
Nonheating0.630.790.590.430.690.830.270.19
Table 5. Uncertainty analysis of PM emissions from wind erosion sources in Southern Xinjiang.
Table 5. Uncertainty analysis of PM emissions from wind erosion sources in Southern Xinjiang.
ParticlesPeriodMean2.5%97.5%Uncertainty
PM10Sand12.143.6620.46(−70%, 68%)
Heating3.340.987.88(−71%, 136%)
Nonheating3.721.037.99(−72%, 115%)
Whole year15.155.2031.27(−66%, 106%)
PM2.5Sand1.670.413.82(−76%, 127%)
Heating0.620.161.34(−74%, 115%)
Nonheating0.990.282.29(−72%, 132%)
Whole year1.830.453.82(−75%, 108%)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gao, S.; Liu, Y.; Zhang, J.; Yu, J.; Chen, L.; Sun, Y.; Mao, J.; Zhang, H.; Ma, Z.; Yang, W.; et al. Soil-Derived Dust PM10 and PM2.5 Fractions in Southern Xinjiang, China, Using an Artificial Neural Network Model. Atmosphere 2023, 14, 1644. https://doi.org/10.3390/atmos14111644

AMA Style

Gao S, Liu Y, Zhang J, Yu J, Chen L, Sun Y, Mao J, Zhang H, Ma Z, Yang W, et al. Soil-Derived Dust PM10 and PM2.5 Fractions in Southern Xinjiang, China, Using an Artificial Neural Network Model. Atmosphere. 2023; 14(11):1644. https://doi.org/10.3390/atmos14111644

Chicago/Turabian Style

Gao, Shuang, Yaxin Liu, Jieqiong Zhang, Jie Yu, Li Chen, Yanling Sun, Jian Mao, Hui Zhang, Zhenxing Ma, Wen Yang, and et al. 2023. "Soil-Derived Dust PM10 and PM2.5 Fractions in Southern Xinjiang, China, Using an Artificial Neural Network Model" Atmosphere 14, no. 11: 1644. https://doi.org/10.3390/atmos14111644

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