Next Article in Journal
Performance Assessment of the COMET Cloud Fractional Cover Climatology across Meteosat Generations
Next Article in Special Issue
Characteristic and Driving Factors of Aerosol Optical Depth over Mainland China during 1980–2017
Previous Article in Journal
Combining TerraSAR-X and Landsat Images for Emergency Response in Urban Environments
Previous Article in Special Issue
Preliminary Investigation of a New AHI Aerosol Optical Depth (AOD) Retrieval Algorithm and Evaluation with Multiple Source AOD Measurements in China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Correcting Measurement Error in Satellite Aerosol Optical Depth with Machine Learning for Modeling PM2.5 in the Northeastern USA

1
Department of Environmental Medicine and Public Health, Icahn School of Medicine at Mount Sinai, New York, NY 10029, USA
2
Department of Geography and Environmental Development, Ben-Gurion University of the Negev, Beer Sheva 8410501, Israel
3
National Aeronautics and Space Administration (NASA) Goddard Space Flight Center (GSFC), Greenbelt, MD 20771, USA
*
Author to whom correspondence should be addressed.
Remote Sens. 2018, 10(5), 803; https://doi.org/10.3390/rs10050803
Submission received: 3 April 2018 / Revised: 11 May 2018 / Accepted: 17 May 2018 / Published: 22 May 2018
(This article belongs to the Special Issue Aerosol Remote Sensing)

Abstract

:
Satellite-derived estimates of aerosol optical depth (AOD) are key predictors in particulate air pollution models. The multi-step retrieval algorithms that estimate AOD also produce quality control variables but these have not been systematically used to address the measurement error in AOD. We compare three machine-learning methods: random forests, gradient boosting, and extreme gradient boosting (XGBoost) to characterize and correct measurement error in the Multi-Angle Implementation of Atmospheric Correction (MAIAC) 1 × 1 km AOD product for Aqua and Terra satellites across the Northeastern/Mid-Atlantic USA versus collocated measures from 79 ground-based AERONET stations over 14 years. Models included 52 quality control, land use, meteorology, and spatially-derived features. Variable importance measures suggest relative azimuth, AOD uncertainty, and the AOD difference in 30–210 km moving windows are among the most important features for predicting measurement error. XGBoost outperformed the other machine-learning approaches, decreasing the root mean squared error in withheld testing data by 43% and 44% for Aqua and Terra. After correction using XGBoost, the correlation of collocated AOD and daily PM2.5 monitors across the region increased by 10 and 9 percentage points for Aqua and Terra. We demonstrate how machine learning with quality control and spatial features substantially improves satellite-derived AOD products for air pollution modeling.

Graphical Abstract

1. Introduction

A useful public health application of satellite remote sensing is to augment sparse monitoring networks and cover large time and space domains when modeling particulate matter for epidemiologic health studies [1]. Recent refinements in remote sensing algorithms have resulted in higher resolution products such as the 1 × 1 km resolution Multi-Angle Implementation of Atmospheric Correction (MAIAC) retrieval algorithm estimating the Aerosol Optical Depth (AOD) as a measure of the density of light scattering particles in the atmospheric column [2,3]. The MAIAC product, derived for the Moderate Resolution Imaging Spectroradiometer (MODIS) instruments, like earlier lower spatial resolution AOD products (e.g., 10 km × 10 km Deep Blue and Dark Target retrieval algorithms), is a key predictor in leading statistical models estimating PM2.5 at the ground level [4,5,6].
Because of the challenge of estimating valid AOD measures over heterogeneous landscapes with varying remote sensing characteristics (e.g., view geometry at acquisition as measured by the relative azimuth), the resulting products can contain patterns that appear anomalous in visualizations, which suggests room for improvement. While previous work has compared the agreement of MAIAC with earlier MODIS 10 × 10 km AOD products and ground monitoring data in different regions and seasons by stratifying the dataset [7], little work has been done to comprehensively understand and correct for measurement error in the MAIAC AOD product.
Quantifying and correcting measurement error in the MAIAC AOD product requires a comparison with a reliable validation dataset. The AErosol RObotic NETwork (AERONET) is a standardized ground-based remote sensing network for measuring aerosol optical depth with a cloud-screened and quality assured data record that is frequently used as a validation for satellite-based AOD products [8].
In this application, we propose and compare three related ensemble machine-learning modeling approaches with a wide range of predictors related to data quality, context, and relevant spatial characteristics. These predictors are used to partially correct measurement error in satellite AOD versus data from AERONET stations across the region which are used as a validation. We demonstrate that this approach improves the MAIAC AOD product over the Northeastern USA. The resulting corrected AOD has a substantially improved correlation with ground-level PM2.5 and thus will be a key predictor in the next generation of satellite-hybrid PM2.5 air pollution models.

2. Materials and Methods

The study region was the Northeast and Mid-Atlantic USA including 13 states from Maine to Virginia. This region includes 629,729 centroids from a fixed grid that is approximately 1 × 1 km in resolution as produced for the MAIAC algorithm (Figure 1). Satellite-derived AOD products from the MAIAC algorithm for both MODIS instruments on the Terra and Aqua satellites were obtained from NASA (version downloaded 16 October 2016). These data include AOD estimates as well as auxiliary variables such as uncertainty estimates, relative azimuth, and additional QC flags such as cloud adjacency masks (full variable list in results). The AOD data were collected from 24 February 2000 to 6 August 2016 and from 4 July 2002 to 31 August 2016 for Terra and Aqua, respectively. The MAIAC dataset is organized by orbit, with the number of acquisitions per satellite per day ranging from 1 to 3; the percentage of days with 2 acquisitions was equal to 78.1% for Terra (with local average overpass times of 9:52 a.m. and 11:31 a.m.) and 82.8% for Aqua (with local average overpass times of 11:51 a.m. and 1:29 p.m.). For those centroids with more than one value of AOD per day (coming from different acquisition times on the same day), making up almost 10% of the data; we kept the record with the lowest AOD uncertainty estimate from the MAIAC dataset.
All global Aerosol Optical Thickness (AOT—a synonymous term for AOD used by AERONET) measurements from AERONET sun photometers were downloaded from https://aeronet.gsfc.nasa.gov/ (accessed 29 March 2017; Level 2.0, cloud-screened and quality-assured data; Version 2.0 Direct Sun Algorithm) and subset to the 79 AERONET stations in the study region (Maine to Virginia) with available data between 2000 and 2015 (Figure 1). AERONET measures were joined to the Aqua or Terra derived MAIAC AOD, when available, from the fixed grid cell centroid closest to the AERONET location, using the AERONET measure closest in time to the satellite overpass (within 60 min).
Our outcome of interest was the difference between AOD and AOT (calculated as AOD-AOT); a residual that approximates the satellite product measurement error. We selected this as our parameter for three reasons: (1) the difference has an easier interpretation because our goal in modeling measurement error was to minimize the difference from the reference AOT; (2) the difference had an approximately normal distribution; and (3) because some ensemble modeling methods resample subsets of covariates, estimating AOT without first subtracting from AOD would lead to regression trees within the ensemble which do not include the AOD as a predictor.
Our modeling approach to estimate measurement error used a set of 52 total predictor variables. These included quality control features that are part of the MAIAC dataset (e.g., relative azimuth at acquisition and the MAIAC algorithm’s own AOD uncertainty estimate), GIS-derived land use and meteorologic covariates (e.g., nearest air temperature from the NOAA reanalysis and proportion of forest within 1 km from the National Land Cover Dataset [9]), and spatially derived covariates engineered to capture characteristics of the data that were observed in visualizations and assigned to each centroid (e.g., number of non-missing AOD centroids within moving windows, difference of AOD from the mean within moving windows of varying edge lengths, and clump size of the number of contiguous non-missing AOD measures within a day). A detailed definition of the model equation including the 52 predictor variables, data sources, and their derivation are included in Appendix A.
In this analysis we trained ensemble machine-learning methods for regression that operate by constructing a multitude of decision trees in a training set and using them to make predictions on a withheld test set. Specifically, we fit three machine-learning algorithms: Random Forest (RF) and two implementations of Gradient Boosting (GB) models. RF uses an ensemble of unpruned decision trees, each grown using a bootstrap sample of the training data, and randomly selected subsets of predictor variables as candidates for splitting tree nodes. The RF prediction for a new observation is the average of the output over all trees. Unlike RF that trains each tree independently, GB grows each tree on the residuals of the previous tree. This means that at each particular iteration, a new weak, base-learner model is trained with respect to the error of the whole ensemble learnt so far. Prediction is accomplished by weighting the ensemble outputs of all the regression trees.
We implemented RF using the randomForest R package [10]. Hyper-parameters were set to use 10,000 trees (ntrees) and to subsample ⅓ of the covariates in each tree (mtry). GB was implemented using two different R functions: gbm (Generalized Boosted regression Models) and xgboost (eXtreme Gradient Boosting) from the gbm and xgboost packages, respectively [11,12]. Although both GBM and XGBoost follow the principle of gradient boosting, XGBoost has some additional changes to improve predictive performance that makes it more of a hybrid of the GB and RF approaches [13]. Specifically, XGBoost uses a more regularized model formalization to control for overfitting, and can also use a random subset of predictor variables at each node like in RF. Hyper-parameters for the GBM were set to use 10,000 trees (n.trees), allow up to 6 splits per tree (interaction.depth) and a learning rate of 0.002 (shrinkage), all selected based on previous tuning experience. For XGBoost, we also applied 5-fold cross-validation to the training set to tune the hyper-parameters of the model. Root Mean Square Error (RMSE) was used to select the optimal model hyper-parameters using the smallest mean value across the 5 folds. The final hyper-parameters for XGBoost for Aqua were 10,000 trees (ntree), allowing up to 5 splits per tree (max_depth), a learning rate of 0.01 (eta), using all features in each tree (colsample_bytree), and using half of the data in each tree (subsample). For Terra the hyper-parameters were the same except for allowing up to 6 splits per tree (max_depth) and subsampling only ⅓ of the covariates in each tree (colsample_bytree). In order to train and validate the performance of these three methods, we split the two datasets (for Terra and Aqua satellites) into training and testing sets. Because the relationship between AOD and ground conditions varies daily [4], the testing sets were created by withholding all observations from randomly selected days across the study period such that the number of withheld testing observations were 15% of the entire dataset. Root Mean Square Prediction Error (RMSPE) of the testing set was used to assess and compare the performance of the three approaches for each satellite.
While predictors can have complex relationships in ensemble models, their contribution can be summarized with variable importance measures to quantify and partial dependence plots to visualize the way that predictions (in our case, measurement error) depends on covariates. Variable importance for the XGBoost model was quantified using the xgb.importance function from the xgboost R package that quantifies how splitting on each feature improves the purity of each node, which in regression tree models is the maximum likelihood estimator of the variance within the node. R functions for permutation testing were used to assess variable importance for both RF and GBM approaches. Both these methods randomly permute each predictor one variable at a time and compute the associated reduction in predictive performance: the higher the reduction, the more important the variable is in predicting the outcome. The only difference in these two functions is that, while GBM permutes the entire training dataset, RF uses only the out-of-bag observations. We also used the R function max.subtree from the randomForestSRC package [14] to compute the first order depth of each variable using RF models—this represents the average number of splits within a tree between the root node and the first split on that variable. The smaller the first order depth the greater the impact of that variable on prediction. For visualization, partial dependency plots show the effect of a predictor variable on the target outcome, after accounting for the average effects of the other predictors—partialling them out.
We also tried fitting XGBoost using just the top 10 or top 20 most important predictors from the XGBoost model to assess the loss of information when using a more parsimonious feature set. In both cases, 5-fold cross-validation was again used on the training set to tune the parameters of the more parsimonious models before assessment on the testing set.
We implemented two methods to estimate the predictive uncertainty in our measurement error model: bootstrapping the ensemble learner (resampling with replacement and refitting the entire model multiple times), and running the Infinitesimal Jackknife (IJ) which has been developed for Random Forest models [15,16].
In previous work, we have demonstrated that AOD is the best single predictor for models to estimate PM2.5 at the ground level, although the complex relationship is improved with daily calibrations [4,17,18]. To demonstrate that the measurement error correction provided by our approach will help with the subsequent modeling of ground-level air pollution, we also compared the raw and corrected values of AOD with PM2.5 measurements from all available monitors within the study region based on the EPA and IMPROVE monitoring networks [4]. The AOD and PM2.5 were compared with the Pearson correlation coefficient using non-missing AOD from the closest grid centroid (within 1 km of the monitoring station) and the daily average PM2.5 concentrations. We remove one monitor (420030064, near Pittsburgh PA) due to aberrant values likely related to the proximity of large industrial facilities, including the Clairton Coke Works. This station is located in the Monongahela river valley across from and <3 km to the Clairton Coke Works, the largest coke manufacturing plant baking coal for steelmaking in the United States.
The resulting dataset included 362 and 381 PM2.5 monitors with 105,798 and 131,788 daily observations with concurrent AOD for Aqua and Terra, respectively. The Pearson correlation is calculated using both the predictions from the measurement error model and using multiple overimputed versions after bootstrapping to account for the predictive uncertainty in our measurement error model [19]. An improvement in the correlation between AOD values and ground-level PM2.5 supports that the corrected AOD would improve future PM2.5 modeling efforts.
All analyses were conducted in R version 3.4.1.

3. Results

The measurement error correction datasets with collocated AERONET AOT and MAIAC AOD used in this analysis included 8531 and 10,278 observations at the AERONET sites for Aqua and Terra, respectively. Table 1 shows that the distribution of AOD from Aqua in this measurement error correction dataset was similar to overall AOD in the region (shown for comparison are selected quantiles from all nearly 50 million non-missing AOD observations for Aqua from 2008 for the full study region).
The Pearson correlation between AOD and AOT in the entire measurement error dataset was 0.86 and 0.89 for Aqua and Terra, respectively. The difference between AOD and AOT (AOD-AOT) had mean and range equal to −0.02 (−0.91, 0.49) and −0.01 (−0.61, 0.66) for Aqua and Terra, respectively (Figure 2).
As described in Materials and Methods, we used three different ensemble learning methods to predict AOT starting from AOD. All models were trained on the training set and their performances were validated by computing the RMSPE on the testing set. Table 2 shows the values of the RMSPE and R2 for models predicting the parameter AOD-AOT in the testing sets for both Aqua and Terra, although the R2 is not directly comparable between the Aqua and Terra datasets because they were built on different datasets and thus are not nested models. For all models, the value of the RMSPE was much lower (up to ~43%, with the best performance from XGBoost) than the root mean square difference between AOD and AOT (0.074 and 0.079 for Aqua and Terra, respectively).
Scatterplots showing the agreement of the XGBoost-corrected data versus the original MAIAC values on the testing set are shown in Figure S1. There is still a strong agreement of the raw AOD and the corrected AOD after applying the measurement error prediction, suggesting that the changes are not drastic with the majority of predictions remaining very similar to their original AOD values. When the XGBoost predicted measurement error model was applied to correct all nearly 50 million non-missing AOD observations for 2008 on Aqua in the full study region, the median absolute difference between the original and corrected AOD was 0.038 and only two percent of observations had an absolute change greater than the interquartile range of the AOD dataset (0.146).
As a visual example of the impact of applying this approach to MAIAC AOD, we generated maps for 16 January 2008 (a testing set day not used in model training) showing the non-missing AOD from the southwestern portion of the region before and after applying the XGBoost model correction (Figure 3). The correction pulls down the right tail of the AOD distribution leading to greater homogeneity in this scene and particularly attenuates some of the higher values seen in small clusters on the left or close to edges that are likely near clouds.
Although all three ensemble approaches use different measures of variable importance, there was generally high agreement in the rank of the variable importance (Figure S2), with several features showing up among the most important in all three modeling approaches for both Aqua and Terra: relative azimuth, AOT uncertainty, long term time trend, windowed differences of AOD over smaller to intermediate scales of 30 km to 210 km, and column water vapor (Figure 4). In general, there was good agreement in the rank of the most and least important variables including the lowest importance for MAIAC cloud mask, which arises infrequently in the data collocated with AERONET measures, and the MAIAC aerosol model flag for dust-affected values. Both of these two variables have little variation in the measurement error dataset over this particular study region. There was less consistency in variable importance rank between modeling approaches for the majority of variables assigned a more intermediate importance.
Partial dependence plots were used to visualize the relation of the most important features with the measurement error parameter (Figure 5 and Figure S3–S5).
To assess whether a simpler model (using fewer features) could achieve similar prediction performance, we also re-fit XGBoost on the Aqua dataset using just the top 10 or top 20 most important predictors from the full XGBoost model. The RMSPE of the testing set was equal to 0.050 (19% worse than the full model) and 0.047 (12% worse than the full model) for models fit with only the top 10 and 20 covariates, respectively.
To reflect predictive uncertainty in our machine-learning models, in addition to predicting corrected AOD, we generated multiple imputation datasets for corrected AOD. This was done both by bootstrapping the original training dataset (RF and XGBoost) or applying the infinitesimal jackknife (IJ) method to estimate variances for each prediction from the RF model only. In all three methods, the variance of the predictions was larger when the absolute difference between AOD and AOT was larger (data not shown).
Because AOD is used as an important predictor in pollution models that estimate ground-level PM2.5, the raw and corrected AOD were correlated with PM2.5 across a network of ground monitoring stations independent of the AERONET AOT. The Pearson correlation between PM2.5 and raw MAIAC AOD was equal to 0.47 and 0.56 for Aqua and Terra, respectively. After correcting the MAIAC AOD using our XGBoost model, the correlations went up to 0.57 and 0.65 for Aqua and Terra, respectively (Table 3). Using Rubin’s rule to combine multiple imputations after a z-transformation [20], the resulting point estimates were the same, with the mean of the correlation between PM2.5 and 5 imputed versions of the XGBoost predicted values of AOT equal to 0.57 (sd = 0.003) and 0.65 (sd = 0.002) for Aqua and Terra, respectively.

4. Discussion

There is a growing record of remote sensing products from an increasing number of sensors, including forthcoming high-resolution sub-hourly coverage from geostationary platforms such as Himawari 8 and GOES-16 [21,22]. The volume of such data makes it possible to construct complex exposure models covering large regions, but also makes the cleaning of these data a challenge with an overwhelming volume of data to visualize and an increasing number of quality metrics to integrate.
We present a machine-learning approach to address measurement error in MAIAC retrievals, a leading AOD product, that makes data cleaning scalable, even when addressing large regions and many years of data. An advantage of adding a measurement error model is that a corrected value can be output as a prediction without reducing the size of the dataset, as is often done when excluding data points with problematic quality control parameters. Thus, this approach can be applied in a two-stage modeling framework where the corrected AOD value is available for further applications such as air pollution modeling or emissions inventories. Using a measurement error framework to correct AOD in this way is novel and differs from previous applications in which machine-learning methods have used uncorrected AOD products as predictors in estimating ground-level air pollutants without first addressing measurement error in AOD [23,24].
When predictions of AOD-AOT are evaluated on a withheld testing set, all three ensemble learning approaches improve on the MAIAC AOD with lower RMSPE for both Aqua and Terra. In both datasets, as expected, the XGBoost model which includes important features of both the Random Forest (random feature subsampling) and the gradient boosting approach outperforms the GBM, which outperforms the RF, in terms of lower RMSPE and higher R2. The additional performance of the gradient boosting approaches GBM and XGBoost over the simpler RF approach requires additional model hyperparameters related to the learning rate and feature subsampling (XGBoost only), but our model tuning suggests that the improved performance of the XGBoost may be due to those additional model characteristics. Comparing raw AOD and predicted AOT with the observed AOT in the testing set, the mode of the AOT is slightly higher and the density is more peaked, although both datasets have a long right tail. The prediction model does not drastically shrink high values as we might have expected if they were due to large systematic errors, such as from cloud contamination as opposed to the long-range transport of smoke from biomass burning, although the training dataset included few truly high AOD values (>1).
While flexible ensemble learning methods are quite performant and lead to excellent predictive performance in data science challenges [25,26], they are sometimes criticized as being less interpretable than a more parsimonious or parametric approach. Variable importance metrics and partial dependence plots can summarize essential relationships and improve the understanding of complex predictive models. For example, the Relative Azimuth (RA) was shown in all three modeling strategies to be one of the most important predictor variables in the variable importance metrics and the partial dependence plot demonstrates that the larger contribution to measurement error occurs when the RA has an angle of >120° in backscattering conditions (which is almost half of the data). This data-driven result is consistent with our expectation of higher error in estimating aerosols in backscattering conditions (when the satellite is oriented between the sun and the earth’s surface) because of a lack of shadows and greater surface brightness that are challenging for aerosol retrieval algorithms [3]. Thus these strategies help to understand how our empirical/statistical findings fit with the physical model underlying the MAIAC retrieval approach and may lead to future contributions in the MAIAC algorithm.
There are multiple methods to estimate variable importance in ensemble learning models. While there was great agreement in the most and least important variables across the three methods we employed, there was some heterogeneity in the rank of the importance of the intermediate variables across methods and with different variable importance metrics. For example, the permutation importance in the RF model, which is known to have difficulty with highly correlated predictors [27], quantified the many different windowed variables (from a highly correlated set of predictors) as much less important than the RF variable importance based on minimal depth, which was in turn highly correlated with the rank importance of the XGBoost model variable importance from node purity (Figure S2).
A major motivation in the feature engineering was to capture aspects of the data that might explain aberrant values seen when plotting individual satellite overpass scenes. For example, the variable “distance to edge”, which measures the distance of a non-missing pixel to the nearest missing pixel, was developed to capture edge effects that might be related to incomplete cloud masking. However, this feature was not as important as variables derived from the anomaly of AOD from moving windows as well as fixed geographic regions, taking advantage of the spatial autocorrelation and generally high homogeneity of AOD measures within a given scene in the Northeastern and Mid-Atlantic USA. Another counter-intuitive finding is that the partial dependence plot for the long-term time trend (operationalized as integer date) suggests that the measurement error in this dataset has been decreasing with time, even though the sensors in the MODIS platform have been aging with an expected decrease in accuracy since their deployment aboard the Aqua and Terra satellites. However, the time trend may be showing patterns in the residual after the removal of the calibration trend from the raw MODIS measurements prior to the application of the MAIAC retrieval algorithm [28].
Given the large number of features considered in these measurement error models, it might be of interest to consider simplifying this approach in future applications by leaving out features that had a low variable importance and that required processing extra data sources. When we ran the XGBoost model with only the top 10 or 20 predictors, there was a moderate decrease in the improvement in measurement error achieved in the testing set. This tradeoff makes sense given that many of the included features had similar and intermediate variable importance measures.
As we propose using a statistical model to update AOD measures, it may be useful to estimate the uncertainty in these predictions as well. While the properties of the Infinitesimal Jackknife (IJ) for estimating predictive uncertainty of estimates from Random Forest models have been previously demonstrated [15], more work is needed to approximate the posterior predictive distribution for other ensemble methods, such as XGBoost. When the same Random Forest model was used to compare the estimated variance of the predictions between bootstrapping and the Infinitesimal Jackknife, we found that the correlation was moderate for variance estimates after only 5 bootstrap resamplings (r = 0.78 excluding a single outlier) and went up further when the variance estimates were computed after resampling 50 times (r = 0.89 excluding a single outlier). A future direction of this work would be to use estimates of the predictive uncertainty in further analyses that employ the corrected AOD values as a predictor of ground-level PM2.5, perhaps by adapting multiple-imputation approaches for measurement error correction [19].
Although our approach to address measurement error in satellite AOD is novel and shows a substantial improvement in the resulting product, it has several limitations. The temporal and spatial coverage of the AERONET dataset used for validation is not representative of the entire space-time domain of interest (e.g., the majority of unique AERONET sites are in urban areas; particularly the DRAGON snapshot campaign in the DC area [29]). Furthermore, because only cloud-screened and quality-assured AERONET station data are used as a reference value for AOD, the most problematic measurements (e.g., when there is a nearby cloud) for satellite-based AOD measures may be underrepresented in our validation dataset making it difficult to train the model to correct these largest outliers. For example, there are few data points in our validation dataset that have very high AOD (AERONET AOT >1.5 only once in Aqua and twice in Terra in our dataset) because true values this large occur only rarely in this region, even though it is not as uncommon in the MAIAC dataset. Given the differences in the overall MAIAC dataset and the subset we use for measurement error correction (collocated with AERONET), future models may be improved by implementing inverse probability weighting of the measurement error dataset back to the originating dataset. Finally, the AERONET measure represents a measurement of AOD at a single point, while the AOD from MAIAC is an estimate over a 1 km2 region, and this can contribute to spatial/temporal misalignment that we have captured within our estimation of measurement error. While our application benefited from a reasonably large number of AERONET stations in the Northeastern USA, this approach has not yet been tested in regions which have few AERONET stations. A future direction for this work will be to examine the generalizability of this approach in other regions (leveraging the global coverage of MAIAC and the AERONET station network) and particularly examine performance where there are fewer unique AERONET stations and collocated observations.
Although extreme (high) values of AOD are relatively rare, these values merit extra attention as they may indicate unusual pollutant scenarios or highly influential outliers if not real. We discovered that for the Aqua satellite, the highest pair of AOD and AOT data points in our collocated measurement error dataset occurred on 10 June 2015 after long range drift of smoke from Canadian wildfires [30]. However, the detection of emitted smoke that has undergone long-range transport is a challenge in the field and may be too infrequent in our measurement error dataset to make a meaningful contribution here.

5. Conclusions

AOD is an important predictor of ground-level fine particulate air pollution (PM2.5) [31]. As a demonstration that adjustment of AOD estimates with our measurement error model reduces noise and enhances the underlying relation with PM2.5 measures, we also compared AOD from Aqua and Terra before and after correction with daily PM2.5 monitors from across the Northeastern US that were not included in any part of our measurement error modeling. We demonstrate that our best measurement error model using XGBoost improves the correlation of collocated MAIAC AOD and daily average PM2.5 by nearly 10 percentage points for both Aqua and Terra. This substantial increase suggests that the use of measurement error corrected MAIAC AOD will be an important advancement for the next generation of satellite-based air pollution models.

Supplementary Materials

The following are available online at https://www.mdpi.com/2072-4292/10/5/803/s1. Supplemental figures show the relation between predictions and target values in the testing dataset (Figure S1), variable importance measures for all three ensemble learning approaches (Figure S2), and partial dependence plots for the effects of AOT uncertainty (Figure S3), column water vapor (Figure S4), and time trend (Figure S5).

Author Contributions

A.C.J. and I.K. conceived and designed the experiments; M.M.D. performed the experiments; A.S. and M.D. contributed in data processing and data analyses; A.L. contributed to interpretation of results and critical discussion of findings. All authors reviewed and contributed to the intellectual content of the manuscript.

Funding

This research was funded by NIH grants UG3 OD023337 and P30 ES023515. A.C.J. was supported by NIH grant R00 ES023450. I.K. was supported by the Israeli Ministry of Science, Space, and Technology.

Acknowledgments

We thank the AERONET federation PIs and their staff for establishing and maintaining the 79 sites used in this investigation. We also gratefully acknowledge the NCEP Reanalysis data provided by the NOAA/OAR/ESRL PSD, Boulder, CO, USA, from their Web site at http://www.esrl.noaa.gov/psd/.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Appendix A.1. Model Specification

The machine-learning ensemble models in this paper used a set of 52 variables that were included because of their hypothesized relation with measurement error in MAIAC AOD or were constructed to capture apparent patterns in visualizations of MAIAC data. The general model specification to estimate the difference between MAIAC AOD and AERONET AOT was:
(aod-aot) ~ air.2m + evap + hpbl + pres.sfc + pr_wtr + shum.2m + uwnd.10m + vis + vwnd.10m + elev + Water_P1km + forestProp_1km + AOT_Uncertainty + Column_WV + RelAZ + maskcloud + maskadj + aerosolmod + distedgekm + sdwin3km + nwin3km + nwin30km + nwin50km + nwin110km + nwin210km + nwin310km + nwin410km + nwin510km + meanwin30km + meanwin50km + meanwin110km + meanwin210km + meanwin310km + meanwin410km + meanwin510km + diffwin30 + diffwin50 + diffwin110 + diffwin210 + diffwin310 + diffwin410 + diffwin510 + percreg10 + percecoreg + meanreg10 + meanecoreg + diffreg10 + diffecoreg + clumpn + clumpmean + dayint + bimon
These variables belong in origin or in meaning within these general groups and are detailed below:
  • AERONET aerosol optical thickness
  • MAIAC variables
  • meteorological and land use variables
  • distance to an edge
  • focal variables
  • regional variables
  • cluster variables
  • temporal variables

Appendix A.2. Detailed Variable Sources and Derivation

Appendix A.2.1. AOD & AOT

As detailed in the Materials and Methods of the manuscript, blue band aerosol optical depth (AOD) came from the MAIAC algorithm variable “Optical_Depth_047”, with fixed grid centroids on a ~1 km2 resolution provided by NASA:
Aerosol optical depth (AOD)
Sun photometer measures of aerosol optical thickness (AOT) were used as ground truth from AERONET stations within the study area (accessed 2017-03-29; Level 2.0, cloud-screened and quality-assured data; Version 2.0 Direct Sun Algorithm):
Aerosol optical thickness (AOT)

Appendix A.2.2. MAIAC Variables

The following variables were extracted without alteration (except where noted) from the MAIAC status_QA HDF formatted files and further details are available in the MAIAC data specification:
  • AOT uncertainty (AOT_Uncertainty)
  • Column water vapor (Column_WV)
  • Relative Azimuth down-sampled from 5 km2 resolution to 1 km2 (RelAZ)
  • Cloud mask (maskcloud)
  • Adjacency mask (maskadj)
  • Aerosol model (aerosolmod)
Two additional variables from the MAIAC quality control data (maskland and clouddetect) were considered and excluded because they took only one value within the measurement error correction dataset.

Appendix A.2.3. Meteorologic Variables

Meteorologic variables were derived from the NCEP North American Regional Reanalysis dataset and the daily average of the nearest measure was assigned to each grid centroid. The included variables, units, and shortened name were as follows:
  • Air Temperature at 2 m (expressed as °C) (air.2m)
  • Accumulated total evaporation (kg/m2) (evap)
  • Planetary boundary layer height (meters) (hpbl)
  • Surface pressure (Pa) (pres.sfc)
  • Precipitable water for the entire atmosphere (kg/m2) (pr_wtr)
  • Specific humidity at 2 m (kg/kg) (shum.2m)
  • U-wind at 10 m (m/s) (uwnd.10m)
  • V-wind at 10 m (m/s) (vwnd.10m)
  • Visibility (meters) (vis)

Appendix A.2.4. Land Use Variables

Elevation was derived from the Shuttle Radar Topography Mission as the average of all 250 m resolution SRTM raster cells within each 1 km grid cell:
Average elevation (elev)
Two land cover variables were derived from the National Land Cover Database 2011 as the percentage of each 1 km grid cell covered by the following land cover categories:
%
Water based on category 11-Open Water (Water_P1km)
%
Forest based on categories 41-Deciduous Forest, 42-Evergreen Forest and 43-Mixed Forest (forestProp_1km)

Appendix A.2.5. Distance to an Edge

Visualizations of MAIAC AOD over the study region seemed to show that high AOD values were often near missing grid cells and therefore these might be falsely elevated related to edge effects near masked regions (e.g., cloud contamination). A new variable was constructed using raster processing as the distance in km to the nearest missing value for each non-missing AOD grid cell:
Distance to edge (distedge)

Appendix A.2.6. Focal Variables

To facilitate contrasts in AOD magnitude and characterization of missingness of AOD, relative to nearby values over various moving windows, a series of additional variables were constructed with raster processing. In addition to the standard deviation (sdwin3km) and number of non-missing (nwin3km) values within a 3 × 3 km window, three main variables (number of non-missing “nwin”, mean of non-missing AOD “meanwin”, and AOD at the centroid minus the mean of the non-missing AOD “diffwin”) were calculated for each grid cell for each day using square moving windows of varying side-lengths (30 km, 50 km, 110 km, 210 km, 310 km, 410 km, 510 km). The focal variable list with variables names was as follows:
  • Standard deviation of non-missing AOD in 3 × 3 km2 window (sdwin3km)
  • # non-missing AOD in 3 × 3 km2 window (nwin3km)
  • # non-missing AOD in 30 × 30 km2 window (nwin30km)
  • # non-missing AOD in 50 × 50 km2 window (nwin50km)
  • # non-missing AOD in 110 × 110 km2 window (nwin110km)
  • # non-missing AOD in 210 × 210 km2 window (nwin210km)
  • # non-missing AOD in 310 × 310 km2 window (nwin310km)
  • # non-missing AOD in 410 × 410 km2 window (nwin410km)
  • # non-missing AOD in 510 × 510 km2 window (nwin510km)
  • AOD mean in 30 × 30 km2 window (meanwin30km)
  • AOD mean in 50 × 50 km2 window (meanwin50km)
  • AOD mean in 1100 × 110 km2 window (meanwin110km)
  • AOD mean in 210 × 210 km2 window (meanwin210km)
  • AOD mean in 310 × 310 km2 window (meanwin310km)
  • AOD mean in 410 × 410 km2 window (meanwin410km)
  • AOD mean in 510 × 510 km2 window (meanwin510km)
  • AOD–AOD mean in 30 × 30 km2 window (diffwin30)
  • AOD–AOD mean in 50 × 50 km2 window (diffwin50)
  • AOD–AOD mean in 110 × 110 km2 window (diffwin110)
  • AOD–AOD mean in 210 × 210 km2 window (diffwin210)
  • AOD–AOD mean in 310 × 310 km2 window (diffwin310)
  • AOD–AOD mean in 410 × 410 km2 window (diffwin410)
  • AOD–AOD mean in 510 × 510 km2 window (diffwin510)

Appendix A.2.7. Regional Variables

Given prior work showing that the calibration of AOD relative to surface conditions can benefit from considering sub-regions within such a large multi-state area, additional variables were constructed that compared each AOD value to the remainder of AOD values within fixed polygons that characterize larger regions. The sets of polygons employed were derived from the Forest Service ecoregions, as well as a simplified set of political boundaries that divide the Northeastern US into ten regions, as previously used in our air pollution modeling. The derived variables included the mean AOD value within a given region, the percentage that each AOD measure was relative to the distribution within the region in which it falls, and the difference between each AOD value and the mean within the region in which it falls. The regional variable list with variable names was as follows:
  • Percentage of political region (percreg10)
  • Percentage of ecoregion (percecoreg)
  • Mean of political region (meanreg10)
  • Mean of ecoregion (meanecoreg)
  • AOD–AOD mean in political region (diffreg10)
  • AOD–AOD mean in political region (diffecoreg)

Appendix A.2.8. Cluster Variables

Because visualizations suggested that contiguous clusters of non-missing AOD seemed to be more correlated than AOD grid cells that were not contiguous (with missing values in between), raster processing was used to clump non-missing AOD values per day with Queen’s adjacency rules and then two summary statistics were calculated for each cluster and assigned to each of the grid cells within that cluster. The cluster variable list with variables names was as follows:
  • Number of non-missing in contiguous cluster (clumpn)
  • Mean of AOD in contiguous cluster (clumpmean)

Appendix A.2.9. Temporal Variables

To allow for temporal trends in measurement error and the other features in the model, two terms were added: an integer date included for long term trends and a six-level indicator variable for seasonal patterns based on collapsing months into bimonthly periods (e.g., Jan/Feb, Mar/Apr). The temporal variable list with variable names was as follows:
  • Integer days since 1970-01-01 (dayint)
  • Bimonthly indicator (bimon)

References

  1. Sorek-Hamer, M.; Just, A.C.; Kloog, I. Satellite remote sensing in epidemiological studies. Curr. Opin. Pediatr. 2016, 28, 228–234. [Google Scholar] [CrossRef] [PubMed]
  2. Lyapustin, A.; Martonchik, J.; Wang, Y.J.; Laszlo, I.; Korkin, S. Multiangle implementation of atmospheric correction (MAIAC): 1. Radiative transfer basis and look-up tables. J. Geophys. Res. Atmos. 2011, 116, 9. [Google Scholar] [CrossRef]
  3. Lyapustin, A.; Wang, Y.; Laszlo, I.; Kahn, R.; Korkin, S.; Remer, L.; Levy, R.; Reid, J.S. Multiangle implementation of atmospheric correction (MAIAC): 2. Aerosol algorithm. J. Geophys. Res. Atmos. 2011, 116. [Google Scholar] [CrossRef]
  4. Kloog, I.; Chudnovsky, A.A.; Just, A.C.; Nordio, F.; Koutrakis, P.; Coull, B.A.; Lyapustin, A.; Wang, Y.; Schwartz, J. A new hybrid spatio-temporal model for estimating daily multi-year PM2.5 concentrations across northeastern USA using high resolution aerosol optical depth data. Atmos. Environ. 2014, 95, 581–590. [Google Scholar] [CrossRef] [PubMed]
  5. Van Donkelaar, A.; Martin, R.V.; Brauer, M.; Hsu, N.C.; Kahn, R.A.; Levy, R.C.; Lyapustin, A.; Sayer, A.M.; Winker, D.M. Global estimates of fine particulate matter using a combined geophysical-statistical method with information from satellites, models, and monitors. Environ. Sci. Technol. 2016, 50, 3762–3772. [Google Scholar] [CrossRef] [PubMed]
  6. Di, Q.; Koutrakis, P.; Schwartz, J. A hybrid prediction model for PM2.5 mass and components using a chemical transport model and land use regression. Atmos. Environ. 2016, 131, 390–399. [Google Scholar] [CrossRef]
  7. Chudnovsky, A.; Tang, C.; Lyapustin, A.; Wang, Y.; Schwartz, J.; Koutrakis, P. A critical assessment of high-resolution aerosol optical depth retrievals for fine particulate matter predictions. Atmos. Chem. Phys. 2013, 13, 10907–10917. [Google Scholar] [CrossRef]
  8. Holben, B.N.; Eck, T.F.; Slutsker, I.; Tanré, D.; Buis, J.P.; Setzer, A.; Vermote, E.; Reagan, J.A.; Kaufman, Y.J.; Nakajima, T.; et al. Aeronet—A federated instrument network and data archive for aerosol characterization. Remote Sens. Environ. 1998, 66, 1–16. [Google Scholar] [CrossRef]
  9. Homer, C.; Dewitz, J.; Yang, L.M.; Jin, S.; Danielson, P.; Xian, G.; Coulston, J.; Herold, N.; Wickham, J.; Megown, K. Completion of the 2011 national land cover database for the conterminous united states—Representing a decade of land cover change information. Photogramm. Eng. Remote Sens. 2015, 81, 345–354. [Google Scholar]
  10. Liaw, A.; Wiener, M. Classification and regression by randomforest. R News 2002, 2, 18–22. [Google Scholar]
  11. Ridgeway, G. Generalized Boosted Regression Models. R Package Version 2.1.3. Available online: https://www.google.com.hk/url?sa=t&rct=j&q=&esrc=s&source=web&cd=1&ved=0ahUKEwihxYfLjpjbAhXMx7wKHao5AHMQFgglMAA&url=https%3A%2F%2Fcran.r-project.org%2Fweb%2Fpackages%2Fgbm%2Fgbm.pdf&usg=AOvVaw0ALtYnS1e_kYe-cOK9ImJD (accessed on 21 May 2018).
  12. Chen, T.; He, T.; Benesty, M.; Khotilovich, V.; Tang, Y. Xgboost: Extreme Gradient Boosting. R Package Version 0.6-4. Available online: cran.fhcrc.org/web/packages/xgboost/vignettes/xgboost.pdf (accessed on 1 January 2017).
  13. Chen, T.; Guestrin, C. Xgboost: A scalable tree boosting system. In Proceedings of the 22nd ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, San Francisco, CA, USA, 13–17 August 2016; ACM: San Francisco, CA, USA, 2016; pp. 785–794. [Google Scholar]
  14. Ishwaran, H.; Kogalur, U.B. Random Forests for Survival, Regression, and Classification (Rf-Src). R Package Version 2.5.0. Available online: https://www.google.com.hk/url?sa=t&rct=j&q=&esrc=s&source=web&cd=1&ved=0ahUKEwilocz_lZjbAhXJU7wKHfp6AQwQFgglMAA&url=https%3A%2F%2Fcran.r-project.org%2Fweb%2Fpackages%2FrandomForestSRC%2FrandomForestSRC.pdf&usg=AOvVaw38a2v6X_POBwVKEC99-EFa (accessed on 21 December 2017).
  15. Wager, S.; Hastie, T.; Efron, B. Confidence intervals for random forests: The jackknife and the infinitesimal jackknife. J. Mach. Learn. Res. 2014, 15, 1625–1651. [Google Scholar] [PubMed]
  16. Brokamp, C. Rfinfer: Inference for Random Forests. R Package Version 0.2.0. Available online: https://github.com/cole-brokamp/RFinfer (accessed on 21 December 2017).
  17. Just, A.C.; Wright, R.O.; Schwartz, J.; Coull, B.A.; Baccarelli, A.A.; Tellez-Rojo, M.M.; Moody, E.; Wang, Y.; Lyapustin, A.; Kloog, I. Using high-resolution satellite aerosol optical depth to estimate daily PM2.5 geographical distribution in mexico city. Environ. Sci. Technol. 2015, 49, 8576–8584. [Google Scholar] [CrossRef] [PubMed]
  18. Kloog, I.; Sorek-Hamer, M.; Lyapustin, A.; Coull, B.; Wang, Y.; Just, A.C.; Schwartz, J.; Broday, D.M. Estimating daily PM2.5 and PM10 across the complex geo-climate region of israel using maiac satellite-based aod data. Atmos. Environ. 2015, 122, 409–416. [Google Scholar] [CrossRef] [PubMed]
  19. Blackwell, M.; Honaker, J.; King, G. A unified approach to measurement error and missing data. Sociol. Methods Res. 2015, 46, 303–341. [Google Scholar] [CrossRef] [Green Version]
  20. Marshall, A.; Altman, D.G.; Holder, R.L.; Royston, P. Combining estimates of interest in prognostic modelling studies after multiple imputation: Current practice and guidelines. BMC Med. Res. Methodol. 2009, 9, 57. [Google Scholar] [CrossRef] [PubMed]
  21. Yumimoto, K.; Nagao, T.M.; Kikuchi, M.; Sekiyama, T.T.; Murakami, H.; Tanaka, T.Y.; Ogi, A.; Irie, H.; Khatri, P.; Okumura, H.; et al. Aerosol data assimilation using data from himawari-8, a next-generation geostationary meteorological satellite. Geophys. Res. Lett. 2016, 43, 5886–5894. [Google Scholar] [CrossRef]
  22. Greenwald, T.J.; Pierce, R.B.; Schaack, T.; Otkin, J.; Rogal, M.; Bah, K.; Lenzen, A.; Nelson, J.; Li, J.; Huang, H.L. Real-time simulation of the goes-r abi for user readiness and product evaluation. Bull. Am. Meteorol. Soc. 2016, 97, 245–261. [Google Scholar] [CrossRef]
  23. Reid, C.E.; Jerrett, M.; Petersen, M.L.; Pfister, G.G.; Morefield, P.E.; Tager, I.B.; Raffuse, S.M.; Balmes, J.R. Spatiotemporal prediction of fine particulate matter during the 2008 northern california wildfires using machine learning. Environ. Sci. Technol. 2015, 49, 3887–3896. [Google Scholar] [CrossRef] [PubMed]
  24. Lary, D.J.; Alavi, A.H.; Gandomi, A.H.; Walker, A.L. Machine learning in geosciences and remote sensing. Geosci. Front. 2016, 7, 3–10. [Google Scholar] [CrossRef]
  25. Chen, T.; He, T. Higgs boson discovery with boosted trees. In Proceedings of the NIPS 2014 Workshop on High-Energy Physics and Machine Learning, Montreal, QC, Canada, 8–13 December 2015; pp. 69–80. [Google Scholar]
  26. Babajide Mustapha, I.; Saeed, F. Bioactive molecule prediction using extreme gradient boosting. Molecules 2016, 21, 983. [Google Scholar] [CrossRef] [PubMed]
  27. Strobl, C.; Boulesteix, A.L.; Kneib, T.; Augustin, T.; Zeileis, A. Conditional variable importance for random forests. BMC Bioinform. 2008, 9, 307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Lyapustin, A.; Wang, Y.; Xiong, X.; Meister, G.; Platnick, S.; Levy, R.; Franz, B.; Korkin, S.; Hilker, T.; Tucker, J.; et al. Scientific impact of modis C5 calibration degradation and C6+ improvements. Atmos. Meas. Tech. 2014, 7, 4353–4365. [Google Scholar] [CrossRef]
  29. Holben, B.; Eck, T.; Schafer, J.; Giles, D.; Sorokin, M. Distributed Regional Aerosol Gridded Observation Networks (Dragon) White Paper. 2011. Available online: http://aeronet.gsfc.nasa.gov/new_web/Documents/DRAGON_White_Paper_A_system_of_experiment.pdf (accessed on 1 August 2017).
  30. NASA Earth Observatory. Smoke over the Mid-Atlantic. Available online: https://earthobservatory.nasa.gov/NaturalHazards/view.php?id=86024 (accessed on 16 September 2017).
  31. Duncan, B.N.; Prados, A.I.; Lamsal, L.N.; Liu, Y.; Streets, D.G.; Gupta, P.; Hilsenrath, E.; Kahn, R.A.; Nielsen, J.E.; Beyersdorf, A.J.; et al. Satellite data of atmospheric pollution for u.S. Air quality applications: Examples of applications, summary of data end-user resources, answers to faqs, and common mistakes to avoid. Atmos. Environ. 2014, 94, 647–662. [Google Scholar] [CrossRef]
Figure 1. Study region in Northeastern and Mid-Atlantic USA with 79 unique AERONET stations showing the number of years of coverage for use in measurement error modeling.
Figure 1. Study region in Northeastern and Mid-Atlantic USA with 79 unique AERONET stations showing the number of years of coverage for use in measurement error modeling.
Remotesensing 10 00803 g001
Figure 2. MAIAC AOD versus AOD-AERONET AOT (aerosol optical thickness) in collocated observations in the Aqua measurement error dataset (n = 8531). Since both AOD and AOT are strictly positive, the empty upper left of the Bland-Altman plot is expected. Marginal histograms show AOD is skewed but AOD-AOT, an estimate of measurement error, is more normally distributed.
Figure 2. MAIAC AOD versus AOD-AERONET AOT (aerosol optical thickness) in collocated observations in the Aqua measurement error dataset (n = 8531). Since both AOD and AOT are strictly positive, the empty upper left of the Bland-Altman plot is expected. Marginal histograms show AOD is skewed but AOD-AOT, an estimate of measurement error, is more normally distributed.
Remotesensing 10 00803 g002
Figure 3. Maps of MAIAC AOD for 2008-01-16 (a) before; and (b) after correction with our XGBoost measurement error prediction model. The inset shows the density of AOD within this scene.
Figure 3. Maps of MAIAC AOD for 2008-01-16 (a) before; and (b) after correction with our XGBoost measurement error prediction model. The inset shows the density of AOD within this scene.
Remotesensing 10 00803 g003
Figure 4. Variable importance predicting measurement error by node impurity from XGBoost for the Aqua and Terra dataset with intervals showing the range of variable importance measures across ten bootstrap-resampling fits of the training dataset.
Figure 4. Variable importance predicting measurement error by node impurity from XGBoost for the Aqua and Terra dataset with intervals showing the range of variable importance measures across ten bootstrap-resampling fits of the training dataset.
Remotesensing 10 00803 g004
Figure 5. Partial dependence plot of measurement error as a function of relative azimuth for the Aqua training set (n = 7280) from the GBM approach. The marginal histogram shows the bimodal distribution of relative azimuth for these Aqua retrievals, with larger errors (further from zero) seen for the second mode with angle >120° in backscattering conditions.
Figure 5. Partial dependence plot of measurement error as a function of relative azimuth for the Aqua training set (n = 7280) from the GBM approach. The marginal histogram shows the bimodal distribution of relative azimuth for these Aqua retrievals, with larger errors (further from zero) seen for the second mode with angle >120° in backscattering conditions.
Remotesensing 10 00803 g005
Table 1. Multi-Angle Implementation of Atmospheric Correction (MAIAC) aerosol optical depth (AOD) from Aqua in the measurement error dataset (collocated with AERONET) and across the region (for 2008).
Table 1. Multi-Angle Implementation of Atmospheric Correction (MAIAC) aerosol optical depth (AOD) from Aqua in the measurement error dataset (collocated with AERONET) and across the region (for 2008).
AODNMeanRange5th25th50th75th95th99th
All AOD measures from 200849,970,0220.1340.000; 4.0000.0190.0410.0960.1870.3890.540
Collocated measurement error dataset85310.1480.019; 1.7360.0210.0530.1050.2000.4320.583
Table 2. Performance predicting AOD-AOT on a test set.
Table 2. Performance predicting AOD-AOT on a test set.
ModelAqua (n = 1251)Terra (n = 1478)
RMSPER2RMSPER2
Raw data (AOD vs. AOT) 10.074N/A0.079N/A
RF0.0470.590.0490.62
GBM0.0440.640.0470.65
XGBoost0.0420.670.0440.68
1 Raw data reports the comparable root mean square difference between raw AOD and AOT.
Table 3. Correlations between PM2.5 and the predicted value of AOT.
Table 3. Correlations between PM2.5 and the predicted value of AOT.
ModelAqua (n = 105,798)Terra (n = 131,788)
Raw MAIAC data0.4730.557
RF adjusted0.5480.633
GBM adjusted0.5670.645
XGBoost adjusted0.5720.649

Share and Cite

MDPI and ACS Style

Just, A.C.; De Carli, M.M.; Shtein, A.; Dorman, M.; Lyapustin, A.; Kloog, I. Correcting Measurement Error in Satellite Aerosol Optical Depth with Machine Learning for Modeling PM2.5 in the Northeastern USA. Remote Sens. 2018, 10, 803. https://doi.org/10.3390/rs10050803

AMA Style

Just AC, De Carli MM, Shtein A, Dorman M, Lyapustin A, Kloog I. Correcting Measurement Error in Satellite Aerosol Optical Depth with Machine Learning for Modeling PM2.5 in the Northeastern USA. Remote Sensing. 2018; 10(5):803. https://doi.org/10.3390/rs10050803

Chicago/Turabian Style

Just, Allan C., Margherita M. De Carli, Alexandra Shtein, Michael Dorman, Alexei Lyapustin, and Itai Kloog. 2018. "Correcting Measurement Error in Satellite Aerosol Optical Depth with Machine Learning for Modeling PM2.5 in the Northeastern USA" Remote Sensing 10, no. 5: 803. https://doi.org/10.3390/rs10050803

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