Next Article in Journal
Monitoring Net Primary Productivity of Vegetation and Analyzing Its Drivers in Support of SDG Indicator 15.3.1: A Case Study of Northeast China
Previous Article in Journal
Providing Fine Temporal and Spatial Resolution Analyses of Airborne Particulate Matter Utilizing Complimentary In Situ IoT Sensor Network and Remote Sensing Approaches
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Refining Spatial and Temporal XCO2 Characteristics Observed by Orbiting Carbon Observatory-2 and Orbiting Carbon Observatory-3 Using Sentinel-5P Tropospheric Monitoring Instrument NO2 Observations in China

1
Key Laboratory of Digital Earth Science, Aerospace Information Research Institute, Chinese Academy of Sciences, Beijing 100094, China
2
International Research Center of Big Data for Sustainable Development Goals, Beijing 100094, China
3
School of Earth and Space Sciences, Peking University, Beijing 100871, China
4
China Highway Engineering Consultants Corporation, Beijing 100089, China
5
Guangxi Key Laboratory of Culture and Tourism Smart Technology, Guilin Tourism University, Guilin 541006, China
*
Author to whom correspondence should be addressed.
Remote Sens. 2024, 16(13), 2456; https://doi.org/10.3390/rs16132456
Submission received: 28 March 2024 / Revised: 27 May 2024 / Accepted: 1 July 2024 / Published: 4 July 2024

Abstract

:
The spatial and temporal variations in the atmospheric CO2 concentrations evidently respond to anthropogenic CO2 emission activities. NO2, a pollutant gas emitted from fossil fuel combustion, comes from the same emission sources as CO2. Exploiting the simultaneous emissions characteristics of NO2 and CO2, we proposed an XCO2 prediction approach to reconstruct XCO2 data based on the data-driven machine learning algorithm using multiple predictors, including satellite observation of atmospheric NO2, to resolve the issue of data gaps in satellite observation of XCO2. The prediction model showed good predictive performance in revealing CO2 concentrations in space and time, with a total deviation of 0.17 ± 1.17 ppm in the cross-validation and 1.03 ± 1.15 ppm compared to ground-based XCO2 measurements. As a result, the introduction of NO2 obtained better improvements in the CO2 concentration responding to the anthropogenic emissions in space. The reconstructed XCO2 data not only filled the gaps but also enhanced the signals of anthropogenic CO2 emissions by using NO2 data, as NO2 strongly responds to anthropogenic CO2 emissions (R2 = 0.92). Moreover, the predicted XCO2 data preferred to correct the abnormally low XCO2 retrievals at satellite observing footprints, where the XCO2_uncertainity field in the OCO-2 and OCO-3 products indicated a larger uncertainty in the inversion algorithm.

1. Introduction

The increase in atmospheric carbon dioxide (CO2) has significantly enhanced global climate warming [1], mainly caused by carbon emissions from human activities [2]. In order to mitigate global warming, countries are adopting measures to reduce and control their anthropogenic carbon emissions. Changes in atmospheric CO2 concentrations can be used as an indicator to monitor the effects of anthropogenic CO2 emission reductions and control [3,4], as well as to evaluate the impacts on natural ecological CO2 emissions due to extreme climate change [5,6]. Thus, accurate data on atmospheric CO2 concentrations in space and time can help us better understand the changing mechanisms of the atmospheric CO2 concentration induced by anthropogenic CO2 emissions and natural effects to support government decision-making policies for carbon emission reduction and control.
Satellite observations of atmospheric CO2 have been used to obtain global CO2 data effectively, with the advantage of global and period coverage. Currently, the XCO2 (columnar average of molar fractions of carbon dioxide in dry air) retrievals derived from CO2 observation satellites such as the Greenhouse Gases Observing Satellite (GOSAT), GOSAT-2, Orbiting Carbon Observatory-2 (OCO-2), and OCO-3 are available from 2009 [7,8,9,10]. These XCO2 retrievals, however, have numerous gaps in space and time due to the satellite observation mode, cloudy or aerosol-contaminated conditions, data quality filtering, etc. These gaps result in missing the fine-tuning changes of CO2 in space and time, making it difficult to correlate the correspondence of CO2 changes with anthropogenic emissions. A data-driven methodology, which is different from the model-based data assimilation method, was developed in this research to fill the gaps and reconstruct satellite-observed XCO2 data.
Applying the geo-statistics of satellite XCO2 retrievals in space and time, some studies have developed a gap-filling method [11,12]. Global spatio-temporal XCO2 datasets in a 1-degree grid over three-day or monthly periods from 2009 have been generated using GOSAT, GOSAT-2, OCO-2, and OCO-3 [13] and have been released in the HARVARD Dataverse [14]. The feasibility and reliability of these datasets have been verified by comparing them with the ground-based measures (The Total Atmospheric Carbon Column Observation Network, TCCON), cross-validation methods, and application analyses to reveal changes in CO2 concentrations corresponding to anthropogenic emissions [15] and extreme climate change. The spatial resolutions of data generated by this method greatly depend on the number of XCO2 retrievals; the more XCO2 retrievals, the higher the spatial resolution. It has been found that this method can only generate data with a resolution of 0.5 degrees when using the available observation data from a single GOSAT, or a combination of GOSAT, OCO-2, and OCO-3, in order to guarantee the accuracy of the filled gaps and the reliability of the generated spatio-temporal data [16].
Recently, machine learning (ML) and deep learning methodologies for fusing multisource data have been developed to reconstruct high-resolution XCO2 data in space and time, driven by the increase in satellite observations that directly or indirectly capture changes in CO2. Algorithms such as Light Gradient-Boosting Machine (LGB), GWNN-GWT, CatBoost (CatB), Random Forest (RF), and Multi-Layer Perceptron (MLP) have been applied to develop XCO2 predictive models using satellite-observed CO2 as the target variable and the multiple parameter data from multisource observations and measurements as predictor variables [17,18,19,20]. The choice of predictor variables strongly influences the accuracy and effectiveness of ML methodology. Previous research has used predictor variables such as NDVI, land use, population, meteorological reanalysis data, and model- simulated CO2 data. However, parameters such as land use and population, suggested to describe the anthropogenic emissions, may not effectively interpret dynamic CO2 emission activity, especially from constantly changing point sources. Furthermore, previous research lacks clarity on the extent to which raw XCO2 retrievals at the footprints are modified after being recomputed based on the ML model and whether the reconstructed XCO2 accurately reflects responses to anthropogenic CO2 emissions.
Atmospheric NO2 concentration data, obtained from observations of Sentinel-5 Precursor/Tropospheric Monitoring Instrument (S5P/TROPOMI), have been widely used in recent years to quantify anthropogenic CO2 emissions, as both NO2 and CO2 emitted by human activity such as fossil fuel combustion originate from the same sources [21,22,23,24]. It has been shown that NO2 observations have the potential to provide much stronger constraints to total annual fossil CO2 emissions [25,26,27]. NO2 emitted from a point source can be sensitively detected with minimal background disturbance, as NO2 is a gas with a short lifetime of only a few hours [28,29,30]. CO2, however, is a long-lived gas [31]. CO2 emissions from sources with weak signals of 1–4 ppm are overshadowed by background levels exceeding 400 ppm, making it difficult to detect enhancements in CO2 emissions with the existing signal-to-noise ratios of sensors in satellites. The magnitude of NO2 columns is generally much larger than background levels around the emitting source, making it a suitable tracer for CO2 emissions and capable of characterizing anthropogenic emissions. Therefore, NO2 can be used as a tracer for CO2 emissions, and previous research has demonstrated the effectiveness of simultaneously using satellite observations of NO2 and XCO2 to detect anthropogenic CO2 emissions [32]. Using NO2 as a proxy for fossil CO2 allows for the quantification of anthropogenic emissions and distinguishes them from biogenic sources of CO2 [33,34].
China is a country with high CO2 emissions and heavy air pollution due to rapid economic development. Anthropogenic CO2 emissions in China are spatially and inhomogeneously distributed, with a significant concentration in the eastern region characterized by high population density, numerous large cities, and industrial enterprises. Evidence of satellite-based CO2 observations strongly correlating with satellite-based NO2 observations have been demonstrated in this area [14]. The Chinese government has been taking measures to reduce carbon emissions and put forward clear time programs for emissions reduction and has outlined clear timelines for emission reduction and control to contribute to the mitigation of global climate warming [35,36,37]. Synergistic analysis of satellite-based CO2 and NO2 could enhance our ability to better monitor the effects of emissions reduction.
Focusing on the effectiveness of predictors and reconstructed XCO2 based on ML, as described above, we introduced NO2 data as one of the predictor variables. Our study area was the Chinese mainland, and we developed an ML-based method to reconstruct XCO2 data from the satellite observations using multiple sources of data. We assessed the performance of reconstructed XCO2 and the contribution and effect of NO2 in reconstructing XCO2 for the quantification of anthropogenic emissions and differentiation from biogenic sources of CO2. Additionally, we analyzed the co-variation of NO2 and CO2 in response to the dramatic reduction in anthropogenic CO2 emissions in a special scenario.

2. Materials and Methods

2.1. Data Used for Reconstructing XCO2

The signals of atmospheric CO2 observed by satellites mainly originate from anthropogenic CO2 emissions, CO2 uptake and emissions (fluxes) from land ecosystems, and atmospheric transmission fluxes over the target area. We collected parameter data significantly related to these sources CO2 [38,39,40,41], along with XCO2 retrievals derived from OCO-2 and OCO-3 observations from January 2019 to December 2022 in the study area, as shown in Table 1.
The XCO2 retrieval data were collected from the OCO-2 XCO2 (OCO-2_L2_Lite_FP 11r) and OCO-3 XCO2 (OCO-3_L2_Lite_FP 10.4r) data products from August 2019 to December 2022. According to the threshold settings of the variables land_fraction < 90 and XCO2_quality_flag = 0 in the raw satellite observation data file, the land area and data quality were filtered, respectively, and only land area and high-quality XCO2 data were used for the analysis. The XCO2 retrievals were derived from CO2 sounder observations on the Orbiting Carbon Observatory-2 (OCO-2) satellite (launched in 2014) and the Orbiting Carbon Observatory-3 (OCO-3) satellite (launched in 2019). The XCO2 data products were generated by the Atmospheric CO2 Observations from Space (ACOS) XCO2 retrieval algorithm, which is an all-physical algorithm and has been filtered through recommended data screening and bias-corrected XCO2 in the Lite files. The retrieval algorithm exhibited a significant regional bias, i.e., the surface pressure was incorrectly estimated in areas of high topographic variability, leading to a more significant bias (~1 ppm) [42]. Cloud-disturbed errors persist in the OCO-2 data, as noted by Massie et al. [43]. The deviation between the bias-corrected XCO2 data and the Total Carbon Column Observing Network (TCCON) is 0.78 ± 1.14 ppm. Additionally, the temporal and spatial coverage of the observation data was not continuous, and there were many blank areas due to limitations of clouds and observation modes.
We collected three-level OFFL NO2 data products in the study area from January 2019 to December 2022 through the GEE platform. The raw NO2 observations were obtained from the Copernicus Ecosystem’s Sentinel-5 Precursor observation satellite, a global atmospheric pollution monitoring satellite launched on 13 October 2017, which carries the Tropospheric Monitoring Raster Spectrometer (TROPOspheric Monitoring Instrument (TROPOMI). Compared with previous remote-sensing instruments for monitoring atmospheric composition, the technical characteristics of the TROPOMI sensor have been greatly improved, with the signal-to-noise ratio increased by a factor of 1 to 5 [44]. The high temporal and spatial resolution and high-precision data observed by TROPOMI can be used to study urban-scale air pollution conditions and air quality monitoring for major events, providing a scientific data basis for policy decisions related to pollution emissions [23,45,46,47]. TROPOMI inverts atmospheric NO2 data using Differential Optical Absorption Spectroscopy (DOAS) in the UV–Vis spectral band. The data processing system is based on the DOMINO-2 product and the Ozone Monitoring Instrument (OMI) EU QA4ECV NO2 reprocessing dataset, and the algorithm is further optimized with the inversion–assimilation–modeling algorithm that uses the TM5-MP chemical transport model at a 1° × 1° resolution in the global three dimensions as the basic element [44]. Since October 2018, the Copernicus ecosystem has provided NO2 data through three Level 2 data products including Near Real-Time (NRTI), Offline (OFFL), and Reprocessing (RPRO). Some of these products may have undergone multiple processing iterations, potentially resulting in improved data quality but delayed availability. Since Level 2 data products are generated on a per-orbit basis, further processing is required for areas where scanned orbits are repeatedly observed. Google Earth Engine (GEE) provides Level 3 data products with a spatial resolution of 1.1132 km, derived from Level 2 data that undergo screening for quality and grid-based reprocessing. The satellite-observed NO2 data are the total vertical column concentration in mol/m2. Since the effect of data observation noise can lead to negative observations in the clean region, only anomalous observations below −0.001 mol/m2 need to be removed, as recommended in the data description document. In this study, monthly-averaged atmospheric NO2 data were calculated for the study area using the GEE platform for online processing, converted to common units of molec/cm2 by a multiplication factor of 6.02214 × 1019, and the spatial resolution was standardized to a 0.01° grid for use in this study.
We collected normalized vegetation index (NDVI) and ECMWF Reanalysis v5 (ERA5) data during the same period as the XCO2 data, which were used as predictors to account for the effects of the vegetation ecological system and atmospheric transport on the atmospheric CO2 concentrations. NDVI, obtained by the Moderate Resolution Imaging Spectroradiometer (MODIS), can account for the CO2 uptake and release of the vegetation ecological system. The seasonal variation in NDVI, caused by vegetation CO2 uptake and release, results in seasonal variations in the CO2 concentration [48,49]. Meteorological data, including T2M (2 m temperature), D2M (dew point temperature), U10 (10 m U-wind fraction), and V10 (10 m V-wind fraction) were obtained from the ERA5 data. These data were derived from the fifth generation of the reanalysis climatic dataset of the European Centre for Medium-Range Weather Forecasts (ECMWF).
Mapping-XCO2 (MXCO2), was used to constrain the global variation in XCO2 as one of the predictors in modeling XCO2 prediction using ML. MXCO2, which represents spatio-temporally continuous XCO2 data, was generated through geo-statistical analysis of the XCO2 retrievals in space and time obtained from multiple satellite observations, including GOSAT (April 2009 to August 2014), OCO-2 (September 2014 to December 2020), and OCO-3 (August 2019 to December 2022). This dataset is available on the HARVARD Dataverse (https://dataverse.harvard.edu/, accessed on 17 August 2021) [14]. The Mapping-XCO2 dataset, which exhibited a −0.29 ± 1.04 ppm deviation compared to the TCCON, enabled examination of temporal and spatial changes at global and regional scales [50,51].
The ground-based XCO2 data released from 2019 to 2022 from Hefei station (117.17°E, 31.9°N) in China, which is in the Total Atmospheric Carbon Column Observation Network (TCCON) [32,52] (https://tccon-wiki.caltech.edu/, accessed on 12 May 2021), were used to validate the accuracy of the reconstructed data by the predictive model. TCCON data, an atmospheric CO2 column concentration derived by ground-based Fourier Transform Spectrometer (FTSE) observations, have been widely used to validate greenhouse gas products from space-based observations.

2.2. Methodology

The modeling XCO2 prediction at spatio-temporal location i can be expressed as follows using the NO2 data in addition to the other parameter variables.
XCO2[i] = f(NO2[i], NDVI[i], T2M[i], D2M[i], U10[i], V10[i], MXCO2[i], T[i])
where f represents the function obtained by the ML algorithm based on the training data. XCO2[i] is the predicted XCO2 by model f at spatio-temporal location i; the variables, including NO2[i], NDVI[i], T2M[i], D2M[i], U10[i], V10[i], MXCO₂[i], and T[i], are the predictors of the spatio-temporal constrain fields for the atmospheric CO2 concentration. Among the predictive parameters, D2M and T2M denote the 2 m dew point temperature and 2 m air temperature, respectively, elucidating the impact of meteorological temperature conditions on the atmospheric CO2 concentration [53]. NO2 serves as an indicator of anthropogenic CO2 emissions, stemming from the same sources [24]. TimeIndex represents timestamps expressed in 1–12 corresponding to the current month, which was used to capture the temporal seasonal variation in the CO2 concentration. NDVI represents physical quantities of the normalized-difference vegetation index, which is used to characterize CO2 uptake and emissions by surface vegetation ecosystems [54]. U10 and V10 represent the physical quantities of the 10 m U-wind component and V-wind component, which are used to characterize the transport effects of the atmospheric wind field [53]. Additionally, MXCO2 represents the spatio-temporal continuum of atmospheric XCO2, serving as a constraint on the overall spatial distribution [14].
The framework for refining XCO2 characteristics in space and time using simultaneously emitted NO2 alongside CO2, as shown in Figure 1, includes ML-based reconstruction of XCO2 data and assessment of the NO2 field constraint to modeling CO2.

2.2.1. Modeling XCO2 Prediction and Reconstructing XCO2 in Space and Time

We built an XCO2 prediction model by co-locating the fp-XCO2 as the Y variable with the predictor variables as the X variable, including NO2, NDVI, T2M, D2M, U10, V10, MXCO2, and TimeIndex in Equation (1).
The training samples used for modeling XCO2 predictions are especially critical, as the type of predictor variables and their spatial scale greatly influence the modeling accuracy and the performance of the predicted XCO2. In addition to incorporating NO2, we applied MXCO2 data generated through mathematical geostatistics of the XCO2 retrievals rather than simulated XCO2 such as CarbonTracker, as carried out by other researchers [17]. This approach aims to constrain the spatio-temporal relationship of XCO2, relying on a data-driven method that identifies a mathematical (constraining) relationship between observations of XCO2 from OCO satellites and multiple parameters. MXCO2 demonstrates the spatio-temporal relationships among the XCO2 retrievals and addresses gaps in satellite observations through mathematical geostatistical methods [13,14].
We assessed the predictive effects of model training data at different scales of a 0.5° grid that is the same as MXCO2 and a 0.1° grid that is the same as the reconstructed XCO2 grid using four ML algorithms, CatB, LGB, MLP, and LSTM, respectively. The results showed that both 0.5° and 0.1° training data without MXCO2 demonstrated much lower prediction accuracy than did those with MXCO2 via cross-validation for all ML algorithms, indicating that MXCO2 can help constrain the spatial relationship of XCO2 during model training. Furthermore, XCO2 predictions from the 0.5°-grid training data demonstrated higher accuracy than those from the 0.1°-grid training data, because resampling MXCO2 to a 0.1° grid within a 0.5° grid could result in the loss of spatio-temporal characteristic information due to the smoothing of scale-up resampling.
As a result of the training data analysis above, we generated a dataset consisting of a set of co-located fp-XCO2 and predictors (NO2, NDVI, T2M, D2M, U10, V10, and MXCO2) from the years 2019 to 2022 in a 0.5° grid/month unit. We then split it into two parts; one is 10% of this dataset, which was randomly extracted as a cross-validation dataset, while the remaining data were used as the training dataset for modeling XCO2 predictions (hereinafter referred to as the T-dataset).
Additionally, we detrended the impact of background on XCO2 (hereafter referred to as dXCO2) by subtracting the monthly average values of the entire XCO2 dataset in the study area for the XCO2 in values in both the T-dataset and P-dataset. This step removed effects from the background due to the atmospheric transportation and the accumulations of CO2 over a long lifetime. Lastly, the XCO2 predictions were calculated for each grid by adding the predicted dXCO2 value to the background trend value used in the same month.
We assessed the performance of the ML algorithms, including CatB, LGB, MLP, and Long Short-Term Memory (LSTM), which have been proven more effective for nonlinear prediction. CatB and LGB are decision tree models, while MLP and LSTM are neural network models. After parameter testing and tuning, the optimal parameter settings for each model are shown below. The CatB model had 10,000 iterations and a learning rate of 0.001. The LGB model had 500 leaf nodes, a model depth of 8, and a learning rate of 0.001. The MLP model had (64, 32, 8) hidden neurons and a learning rate of 0.001. The LSTM model had 50 hidden neurons and a learning rate of 0.001. As a result, the predictions by CatB significantly outperformed those of LGB, MLP, and LSTM through model cross-validation, demonstrating reasonable performance in predicting XCO2 in both space and time. CatB showed the smallest deviation from fp-XCO2, which is 0.17 ± 1.17 ppm (R2 = 0.81), while LGB, MLP, and LSTM showed deviations of 0.21 ± 1.18 ppm (R2 = 0.79), 1.49 ± 3.18 ppm (R2 = 0.47), and 1.03 ± 2.16 ppm (R2 = 0.58), respectively (see Figure A1). Specific comparative validation screenings of the model results are outlined in the Discussion section. Therefore, we applied CatB for modeling XCO2 predictions to reconstruct XCO2.

2.2.2. Evaluation of Reconstructed XCO2 and Effects of NO2 Constraints

The multiple predictor variable data (NO2, NDVI, T2M, D2M, U10, V10, and MXCO2) in Equation (1), each with different spatial and temporal resolutions as shown in Table 1, were integrated into the dataset at a resolution of 0.1° grid/month unit for each variable from the years 2019 to 2022 (hereinafter referred to as the P-dataset). This integration was achieved through resampling, where those variables with a resolution of less than 0.1° in the grid were averaged within each grid for each variable, and MXCO2 was resampled using the nearest-neighbor method.
XCO2 was predicted using the XCO2 prediction model built previously with the P-dataset and reconstructed XCO2 data at a resolution of 0.1°/month from 2019 to 2022, as shown in Figure 1.
The accuracy and performance of the reconstructed XCO2 were implemented through three methods, namely, model cross-validation, ground-based observed XCO2, and analysis of the predictor’s contribution and co-response of NO2 and CO2 to anthropogenic emissions, as well as the effects of NO2 for disentanglement from biogenic sources of CO2.
We extracted ground-based XCO2 data from the TCCON site in Hefei, obtained between 12:00 and 14:00 around the local observation time of OCO series satellites (13:30), and calculated monthly averaged values (hereafter referred to as T-XCO2). The paired data of the co-located predicted XCO2 and T-XCO2 were used to calculate three metrics, the coefficient of determination (R2), the root mean square error (RMSE), and the mean absolute percentage error (MAPE), using the following equations to evaluate the accuracy of the predicted XCO2.
The coefficient of determination, R-squared (R2), was calculated by Equation (2), where a value closer to 1 indicates a better fit of the model.
R 2 = 1 ( Y i Y ^ ) ( Y i Y ¯ )
The formula for the root mean square error (RMSE) and mean absolute prediction error (MAPE) are as follows, respectively:
X R M S E = i = 1 N Y i Y ^ i 2 N
X M A P E = 100 % N i = 1 N | Y i Y ^ i Y |
where Y i is the observed value, Y ^ is the model predicted value, Y ¯ is the mean of the observations, and N is the number of samples.
We applied the SHAP (Shapley Additive exPlanations) method to assess the contribution of predictor variables to the XCO2 predictions and to explore the contribution of NO2 co-emitted with fossil CO2 during combustion, along with the co-variation of NO2 and CO2 in response to anthropogenic CO2 emissions through clustering analysis. Clustering analysis of the spatio-temporal characteristics of NO2 can demonstrate the emission intensity in clustered area units, thereby integrating different chemical processes of NO2 and CO2 in the atmosphere. This approach allows us to investigate the co-variation of NO2 and XCO2 in response to anthropogenic CO2 emissions by statistically analyzing clustered areas, which helps avoid the effects of chemical processes in NO2 and CO2. The clustering analysis of NO2 data at a resolution of 0.1°/month from the years 2019 to 2022 classified the study area into 14 categories using the K-means method.

3. Results

3.1. Model Prediction Accuracy and Performance of the Reconstructed XCO2

3.1.1. Accuracy of the XCO2 Predictions

Figure 2 demonstrates the predicted XCO2 from the CatB-based built predictive model compared to both the satellite-based fp-XCO2 and the ground-based observed XCO2.
The results from the cross-validation, as shown in Figure 2a, demonstrate that the R2, RMSE, and MAPE of the XCO2 predictions compared with the fp-XCO2 are 0.81, 2.26 ppm, and 0.38, respectively, with a total deviation of 0.17 ± 1.17 ppm. The model’s predicted XCO2 tended to be higher than the satellite-observed XCO2 in the low-value range of fp-XCO2, which implies that the model predictions increase those low XCO2 retrievals mostly in the areas without anthropogenic emissions or irregular values. The validation of the predicted XCO2 against T-XCO2, shown in Figure 2b, demonstrates R2, RMSE, and MAPE values of 0.79, 1.52 ppm, and 0.29, respectively. The total deviation of the predicted XCO2 was 1.03 ± 1.15 ppm, similar to the deviation of the co-located fp-XCO2 and TCCON site, which was 0.89 ± 1.23 ppm.

3.1.2. Performance of the Model Predictions

We applied the XCO2 prediction model constructed by the CatB algorithm and T-dataset to generate an XCO2 dataset at a resolution of 0.1° grid/month, denoted as Cm-XCO2, from 2019 to 2022 using the predictor dataset from 2019 to 2022. Figure 3 shows the averaged values for each grid calculated from the model-predicted XCO2, as well as the original satellite-observed XCO2 values within the 0.1° grid. The model-predicted XCO2 exhibited significantly finer spatial features and appeared more reasonable than the original satellite XCO2 retrievals, particularly evident in the resolution of anomalous high XCO2 in the southern region, and the western Taklamakan Desert in the original satellite observations in the study area.
We assessed the impact of model prediction on the original satellite-observed XCO2 (fp-XCO2) that was re-calculated by the predicting model comparing the co-located fp-XCO2 and Cm-XCO2. Figure 4 shows a diagram, histogram, and the difference of the co-located Cm-XCO2 and fp-XCO2, along with a comparison with the posterior uncertainty in the XCO2 algorithm released in L2 product data [55].
It can be found from Figure 4a,b that the model predicting XCO2 narrowed the range of fp-XCO2 values from 399.06–428.67 ppm to 403.04–424.14 ppm within the cumulative frequency range of 5–95%. The low fp-XCO2 values of 395–402 ppm were enlarged to 403–417 ppm using the predictive model, occurring mostly in mountainous areas with highly undulating terrain along the edge of the Tibetan Plateau, as shown in Figure 4c. Here, the differences between Cm-XCO2 and fp-XCO2 ranged from 1 to 4 ppm. It is known that a CO2 concentration of less than 400 ppm is unreasonable since 2019 according to the global CO2 background concentration (see Figure A2), where the minimum of CO2 concentration was 408 ppm in 2019, as observed the from ground-based XCO2 measurements from the TCCON station in Hefei shown in Figure 2b.
Additionally, comparing Figure 4c and Figure 4d reveals that the large differences between Cm-XCO2 and fp-XCO2 correspond to large posterior uncertainty in the XCO2 retrieval data products derived by the OCO L2 algorithm [56] over the edge of the Tibetan Plateau, respectively, which has higher elevation and topographic variations. This is likely because the sharp variations in elevation in the undulating terrain result in anomalous values of fp-XCO2. Currently, the remaining uncertainties in XCO2 retrievals may indicate large biases, especially in regions with large topographic variations [54]. Therefore, the reconstructed XCO2 for these abnormally low values of fp-XCO2, by constraint computation of XCO2 through the ML of the multiple variables, rectified these XCO2 retrievals with large uncertainties, which could be induced by the uncertainty of input parameters in the algorithms of XCO2 retrievals [55].
The high values of fp-XCO2 ranged from 422 to 430 ppm, which are currently unreasonable, as the Global Atmosphere Watch Programme (GAW) global background station located in Waliguan, Qinghai, China, observed that the atmospheric CO2 concentrations peaked at 419.3 ppm in 2022 [57]. These high values are predicted to be reduced to low values ranging from 415 to 420 ppm in Cm-XCO2, mostly in eastern areas with differences ranging from −2 ppm to −3 ppm.

3.2. Co-Variation of CO2 and NO2 to Anthropogenic CO2 Emissions

3.2.1. NO2 Constraints Enhancing the XCO2 Response to Anthropogenic Emissions

The larger the SHAP value of a predictor variable, the greater its contribution to the model prediction. The contribution of an individual variable is ranked in order of contributions by the average SHAP value, as shown in Figure 5. This is defined as the average of the absolute SHAP value (in ppm) and the standard deviation calculated from the training data from 2019 to 2022.
The results of the SHAP values show that D2M, NO2, and TimeIndex significantly contributed to modeling XCO2 predictions much more than the other variables. The largest contributor, D2M, and third-largest, TimeIndex, are likely significant because the distinct seasonal variation in CO2 is strongly impacted by CO2 uptake and release in terrestrial ecosystems, which generally depend on temperature and water vapor, as explained by D2M and monthly time. The contribution of the vegetation index NDVI was low, ranking second from the bottom. The spatial NDVI demonstrated the different seasonal variations in CO2 absorption capacity over various land surfaces with different vegetation and non-vegetation densities. Wind can result in CO2 flowing in space, diminishing the CO2 signals in the atmosphere from CO2 fluxes of land ecosystem surfaces [19].
NO2 ranked second in the contribution of multiple variables to modeling XCO2 prediction. NO2, a pollutant gas emitted from fossil fuel combustion, strongly affects the modeling of XCO2 prediction due to its co-emission with CO2. NO2 showed the strongest linear relationship with Cm-XCO2 (R2 = 0.57) among the predictor variables (see Figure A3), while D2M and NDVI showed a nonlinear relationship with Cm-XCO2. Furthermore, we investigated the co-variation of NO2 and XCO2 in relationship to anthropogenic CO2 emissions.
NO2, as a tracer of CO2 emissions from fossil fuel combustion, can constrain the predictive model to enhance the information on CO2 from the anthropogenic emissions. The spatial distribution of NO2, as shown in Figure 6a, generally aligned with the anthropogenic CO2 emissions (Figure 6b), and the hotspots of emissions demonstrated the spread of high NO2 concentrations. NO2 showed a strong response to the anthropogenic CO2 emissions with an R2 of 0.92, as shown in Figure 6c, when calculating their correlation between NO2 and EDGAR emissions using the averaged values of the grids within each clustered area derived from the spatio-temporal clustering analysis of NO2 data. This result implies that NO2, as a predictor variable, could strengthen the response to anthropogenic CO2 emissions in modeling XCO2 predictions. The predicted XCO2 for those locations of the satellite-observed XCO2 (fp-XCO2) showed an R2 of 0.79 with the co-located anthropogenic emission, which is higher than the R2 of fp-XCO2 (R2 = 0.66). Here, the CO2 response to anthropogenic emissions (66–79%) was lower than that of NO2 (92%), reasonably indicating that atmospheric CO2 still includes natural emissions from land ecosystems and transported CO2 fluxes.
The relationship between the co-located NO2 and fp-XCO2 reached up to 0.74 based on the clustered areas, compared with the calculations based on the grids (R2 = 0.57 in Figure A3b). The clustered areas derived from the spatio-temporal clustering analysis of NO2 data smoothed the impacts of different spreads and lifetimes between NO2 and CO2 in the grid by averaging said grids for each clustered area. The strong correlations between the predicted XCO2 (Cfp-XCO2 and Cm-XCO2) and NO2, with R2 values of 0.86 and 0.92, respectively, as shown in Figure 7, indicated that NO2, as one of the predictors, can constrain XCO2 in modeling XCO2 predictions. This is consistent with Emily G. Yang et al.’s suggestion that the satellite-based NO2 can serve as a proxy species to constrain CO2 [34].

3.2.2. Co-Response of NO2 and CO2 Concentrations to Anthropogenic Emissions under Special Scenarios of Human Activity

The significant reduction in human activities during the 2020–2022 COVID-19 period resulted in decreases in NO2 and CO2 from anthropogenic emissions [14]. We analyzed the co-variation of CO2 and NO2 under this particular reduction in anthropogenic emissions from 2019 to 2022 to find how NO2 and XCO2 synchronized in response to an abnormal reduction in anthropogenic emissions. Based on the clustering results of the spatio-temporal characteristics of the NO2 concentration and anthropogenic CO2 emissions in China, we selected 13 regions of interest (ROIs) with high NO2 values and anthropogenic emissions, as shown in Figure 8a. ROI1–ROI13 (referred to as R1–R13 in Figure 8) indicate Wuhan (R1-Wuhan), Shanghai (R2-Shanghai), the Yangtze River Delta (R3-YRD), Beijing–Tianjin–Hebei (R4-BTH), Xi’an (R5-Xian), Shanxi Coal Mining Belt (R6-ShanxiCoal), Jinan (R7-Jinan), Chengdu–Chongqing (R8-ChengduC), Guangzhou (R9-Guangzhou), the strip west-northwest of Urumqi toward Shihezi (R10-Wulumuqi), Yinchuan (R11-Yinchuan), the strip south of Hohhot (R12-Huhehaote), and Shenyang (R13-Shenyang).
It is known that the maximum anthropogenic emissions generally occur in the winter season when the XCO2 shows the largest enhancements [14]. Therefore, we calculated the enhancement relative to the background value in the winter (December–February) for each ROI (winter value in ROIs minus the winter average of the overall study area in the same year) for the four years from 2019 to 2022 (hereafter referred as to ΔXCO2 and ΔNO2), as shown in Figure 8b. Figure 8b shows that both ΔNO2 and ΔXCO2, after 2019, ahead of the COVID-19 outbreak and control, were lower than the values in the normal year of 2019 from R1 to R7, especially in R1-Wuhan in the eastern part of China, where ΔXCO2 was around 3 ppm. These yearly changes clearly respond to the unexpected reduction of human activity in 2020 and 2021 during the COVID-19 outbreak and control, with human economic activity not returning to 2019 levels in 2022. The reduction in anthropogenic activity during the COVID-19 period resulted in a decrease in ΔNO2 of 0.4–1.0 × 1016 molec/cm2 and in ΔCO2 of 0.5–1.7 ppm compared to 2019. The largest decrease in ΔCO2, 1.7 ppm, occurred when emissions from human activity were minimized in Wuhan.
The ΔCO2 in R8–R13 were less than those of R1–R7, and the interannual variations were also small due to a lesser impact from COVID-19. It should be noted that ΔNO2 in R10-Urumqi presented the largest value among the ROIs abnormally, with the maximum value, 3.4 × 1016 molec/cm2, occurring in 2021, not in the normal year of 2019. The maximum interannual difference was up to 75% during 2019–2022, while ΔCO2 did not present similar changes to ΔNO2 correspondingly, with the maximum interannual difference being only 12%. This abnormal ΔNO2 was due to high emissions from many air-polluting enterprises (mainly the coal and chemical industry and non-ferrous metals) that have been rapidly increasing in recent years in this region, resulting in it having the worst air pollution in China in recent years. The discrepancy, where the change in CO2 did not correspond to the change in NO2, was likely caused by two reasons. One reason is probably due to the fossil fuel combustion processes in these air-polluting enterprises, of coal chemical industry and non-ferrous metal, in which NO2, the pollutant gas emitted by companies, has been rising, while CO2 has reached its maximum. That is, the emitted CO2 does not respond to an increase in emitted NO2 from these sources when NO2 emissions increase beyond a certain level. Another reason is probably due to the uncertainties of XCO2 data; the aerosols, which are one of the key input parameters in the XCO2 retrieval algorithm, likely introduced large biases due to the heavy air pollution (see Figure 4d), even if the ML predictions (see Figure 4c) modified some of their biases.
The extreme reductions in anthropogenic emissions in 2020 and 2021 impacted the corresponding variation in NO2 and XCO2 when comparing NO2 and XCO2 in these two years to the normal year 2019, as shown in Figure 8c. We found that the R2 in 2020 and 2021 (0.49–0.50) was less than in 2022 (0.63) when anthropogenic emissions had begun to recover but not yet returned to 2019 levels. This implies that CO2 may inexactly respond to NO2 changes in these extreme scenarios and during extreme NO2 events like R10-Urumqi.
Further, in contrast to 2019, the monthly differences between monthly values from 2020 to 2022 and the same month in 2019 showed that XCO2 inexactly responded to NO2 as well. Figure 9 shows an example of these differences for R1-Wuhan, R2-Shanghai, R3-YRD, and R4-BTH, where the fluctuations in human economic activities were the highest during the COVID-19 period. Figure 9 shows a co-decrease in NO2 and XCO2 during the first period of COVID-19, covering January–March of 2020, especially in R1-Wuhan, which experienced an extreme reduction in human activities due to COVID-19. NO2 showed a decrease of 1.3 × 1016 molec/cm2 in February 2020, corresponding to a reduction of 1.4 ppm in XCO2. The faint co-decrease in NO2 and XCO2 during the COVID-19 controlling period in April 2022 in R2-Shanghai is shown as well, with a decrease of 0.5 × 1016 molec/cm2 in NO2, corresponding to a reduction of 0.5 ppm in XCO2.
XCO2 still demonstrated a yearly increase of, on average, 2.3 ppm from 2020 to 2022, which implies that regional reductions in anthropogenic emission activities, such as in China, cannot effectively mitigate the increase in CO2 globally. This means that to mitigate further increase in CO2, global action is required.

4. Discussion

Considering that the prediction results of different machine learning algorithms differ, we further compared the prediction performance of LGB, MLP, and LSTM. The results showed that the prediction results of all three models were worse than CatB. The R2 and total deviation of the cross-validation for the four models (CatB, LGB, MLP, and LSTM) were 0.81, 0.79, 0.47, and 0.58, and 0.17 ± 1.17 ppm, 0.21 ± 1.18 ppm, 1.49 ± 3.18 ppm, and 1.03 ± 2.16 ppm (see Figure 2 and Figure A1), respectively. The predictions of CatB and LGB, based on decision trees, were better than those of MLP and LSTM, based on neural networks. LGB and LSTM generally presented overestimates, while MLP presented severe underestimates (see Figure 3 and Figure 10).
These results are likely due to the fact that different machine learning algorithms have different abilities to capture the features of multisource satellite observations. CatB, and LGB are optimized algorithms based on the Gradient-Boosting Decision Tree (GBDT) algorithm, which is the most widely used ML algorithm. GBDT applies weak classifiers (decision trees) for training to obtain an optimal model iteratively, which has the advantages of a good training effect and being hard to overfit. MLP and LSTM are forward-structured artificial neural networks containing an input layer, an output layer, and several hidden layers. The error backward propagation technique was employed to train the parameters of the neural network model and is capable of handling nonlinear separable problems.
The total contribution of the model predictor parameters to the predictions (see Figure 5 and Figure 11) showed that the contribution of the predictor variables in MLP and LSTM were 24.7%, 13.6%, 13.6%, 10.8%, 10.7%, 10.4%, 8.5%, and 7.7% (NDVI, NO2, MXCO2, T2M, V10, U10, D2M, and TimeIndex) and 22.7%, 19%, 16.5%, 11.9%, 11.2%, 8%, 6.3%, and 4.5% (TimeIndex, MXCO2, NO2, D2M, NDVI, T2M, V10, and U10), respectively. It can be seen that NDVI and MXCO2 had unusually high contributions in MLP, as well as TimeIndex and MXCO2 in LSTM, which led to significantly worse predictions in both models. In addition, TimeIndex contributed a higher percentage in LGB and LSTM than in CatB. As described in Section 2.2, for the physical significance and impact of the predicted parameters, overconsideration of the temporal increase in the atmospheric CO2 concentration leads to an overall overestimation of its predictions.

5. Conclusions

The spatio-temporal discontinuities in the original satellite XCO2 observations and the large number of data gaps make it difficult to accurately reveal the spatial and temporal characteristics of the atmospheric CO2 concentration. We reconstructed satellite-observed XCO2 data using multiple predictors, including satellite-observed NO2 with co-emissions from the same sources based on a machine learning algorithm to generate spatially and temporally continuous monthly XCO2 data from January 2019 to December 2022 in Chinese regions. The accuracy of the ML predictive model and the effectiveness of introducing NO2 to enhance the XCO2 response to anthropogenic emission activities were assessed through the cross-validation, validation against TCCON observations, and comparison with original satellite XCO2 retrievals.
The results showed that the predictive model had good predictive performance in terms of revealing the spatial and temporal concentrations of CO2, with an R2 and total deviations of 0.79, 0.17 ± 1.17 ppm, and 0.81, 1.03 ± 1.15 ppm, respectively, compared to ground-based XCO2 measurements and model cross-validation. The reconstructed XCO2 data not only fills in the gaps but also corrects the large uncertainties of the original observations over areas with high elevation and steep terrain. The strong correlation of NO2 with the anthropogenic CO2 emissions (R2 = 0.92), due to simultaneous anthropogenic CO2 emissions, implies that the introduction of NO2 can enhance the anthropogenic emissions information in the model-predicted XCO2 with R2 = 0.76, which is higher than the R2 = 0.66 found for the raw satellite-observed XCO2.
The findings, by analyzing co-response variations of XCO2 and NO2 under special anthropogenic emission scenarios during the COVID-19 control period of 2020–2022 in the study area, indicate a synchronized CO2 and NO2 response to the dramatic reduction in anthropogenic activities; however, CO2 could incompletely respond to NO2 changes in the case of overly heavy gas polluting scenarios. This XCO2 and NO2 co-response variation mechanism to anthropogenic CO2 emissions needs further verification via practical CO2 and NO2 emissions measures.

Author Contributions

Conceptualization, K.G., L.L. and M.S.; Methodology, K.G. and H.S.; Software, K.G., Z.J. and H.S.; Validation, K.G.; Formal analysis, K.G.; Data curation, K.G.; Writing—original draft, K.G.; Writing—review & editing, K.G. and L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program of China (grant no. 2022YFC3800700 and 2020YFA0607503).

Data Availability Statement

The research data presented in this study are available on request from the corresponding author. The data are not publicly available as this project is still in the research phase.

Acknowledgments

We are grateful for the OCO-2 v11r data and OCO-3 v10.4r data, which were provided by the OCO-2/OCO-3 project at the Jet Propulsion Laboratory, California Institute of Technology, and obtained from the OCO-2/OCO-3 data archive maintained at the NASA Goddard Earth Science Data and Information Services Center. We thank the European Space Agency (ESA) and Google Earth Engine for providing Sentinel-S5P NO2 products and the World Data Centre for Greenhouse Gases (WDCGG) for providing global atmospheric CO2 data. We also acknowledge the Land Processes Distributed Active Archive Center (LP DAAC) at the National Aeronautics and Space Administration (NASA) for sharing land cover type and NDVI data derived from MODIS. We thank the Total Carbon Column Observing Network (TCCON) for providing XCH4 observed data products.

Conflicts of Interest

The authors declare no conflicts of interest. Author Mengya Sheng was employed by the company China Highway Engineering Consultants Corporation. The company China Highway Engineering Consultants Corporation had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Appendix A

Figure A1. Plot of cross-validation results for models, (a) LGB, (b) MLP, and (c) LSTM.
Figure A1. Plot of cross-validation results for models, (a) LGB, (b) MLP, and (c) LSTM.
Remotesensing 16 02456 g0a1
Figure A2. Time series of monthly mean XCO2 values from CatB and satellite raw observations.
Figure A2. Time series of monthly mean XCO2 values from CatB and satellite raw observations.
Remotesensing 16 02456 g0a2
Figure A3. CatB model predictions versus parameters (D2M, NO2, T2M, MXCO2, and NDVI on the x-axis of (ae), respectively).
Figure A3. CatB model predictions versus parameters (D2M, NO2, T2M, MXCO2, and NDVI on the x-axis of (ae), respectively).
Remotesensing 16 02456 g0a3

References

  1. Wang, H.; Liu, Z.; Zhang, Y.; Yu, Z.; Chen, C. Impact of Different Urban Canopy Models on Air Quality Simulation in Chengdu, Southwestern China. Atmos. Environ. 2021, 267, 118775. [Google Scholar] [CrossRef]
  2. Hsueh, Y.-H.; Li, K.-F.; Lin, L.-C.; Bhattacharya, S.K.; Laskar, A.H.; Liang, M.-C. East Asian CO2 Level Change Caused by Pacific Decadal Oscillation. Remote Sens. Environ. 2021, 264, 112624. [Google Scholar] [CrossRef]
  3. Liang, M.; Zhang, Y.; Ma, Q.; Yu, D.; Chen, X.; Cohen, J.B. Dramatic Decline of Observed Atmospheric CO2 and CH4 during the COVID-19 Lockdown over the Yangtze River Delta of China. J. Environ. Sci. 2023, 124, 712–722. [Google Scholar] [CrossRef]
  4. Fang, S.; Du, R.; Qi, B.; Ma, Q.; Zhang, G.; Chen, B.; Li, J. Variation of Carbon Dioxide Mole Fraction at a Typical Urban Area in the Yangtze River Delta, China. Atmos. Res. 2022, 265, 105884. [Google Scholar] [CrossRef]
  5. Fu, Y.; Sun, W.; Luo, F.; Zhang, Y.; Zhang, X. Variation Patterns and Driving Factors of Regional Atmospheric CO2 Anomalies in China. Environ. Sci. Pollut. Res. 2022, 29, 19390–19403. [Google Scholar] [CrossRef]
  6. Wang, W.; He, J.; Feng, H.; Jin, Z. High-Coverage Reconstruction of XCO2 Using Multisource Satellite Remote Sensing Data in Beijing–Tianjin–Hebei Region. Int. J. Environ. Res. Public Health 2022, 19, 10853. [Google Scholar] [CrossRef]
  7. Crisp, D.; Fisher, B.M.; O’Dell, C.; Frankenberg, C.; Basilio, R.; Bösch, H.; Brown, L.R.; Castano, R.; Connor, B.; Deutscher, N.M.; et al. The ACOS CO2 Retrieval Algorithm–Part II: Global XCO2 Data Characterization. Atmos. Meas. Tech. 2012, 5, 687–707. [Google Scholar] [CrossRef]
  8. Nakajima, M.; Kuze, A.; Suto, H. The Current Status of GOSAT and the Concept of GOSAT-2. In Proceedings of the Sensors, Systems, and Next-Generation Satellites XVI, Edinburgh, UK, 24–27 September 2012; Meynart, R., Neeck, S.P., Shimoda, H., Eds.; SPIE: Bellingham, WA, USA, 2012; p. 853306. [Google Scholar]
  9. Kataoka, F.; Crisp, D.; Taylor, T.; O’Dell, C.; Kuze, A.; Shiomi, K.; Suto, H.; Bruegge, C.; Schwandner, F.; Rosenberg, R.; et al. The Cross-Calibration of Spectral Radiances and Cross-Validation of CO2 Estimates from GOSAT and OCO-2. Remote Sens. 2017, 9, 1158. [Google Scholar] [CrossRef]
  10. Jin, C.; Xue, Y.; Jiang, X.; Zhao, L.; Yuan, T.; Sun, Y.; Wu, S.; Wang, X. A Long-Term Global XCO2 Dataset: Ensemble of Satellite Products. Atmos. Res. 2022, 279, 106385. [Google Scholar] [CrossRef]
  11. Guo, X.; Zhang, Z.; Cai, Z.; Wang, L.; Gu, Z.; Xu, Y.; Zhao, J. Analysis of the Spatial–Temporal Distribution Characteristics of NO2 and Their Influencing Factors in the Yangtze River Delta Based on Sentinel-5P Satellite Data. Atmosphere 2022, 13, 1923. [Google Scholar] [CrossRef]
  12. Zhang, L.; Li, T.; Wu, J. Deriving Gapless CO2 Concentrations Using a Geographically Weighted Neural Network: China, 2014–2020. Int. J. Appl. Earth Obs. Geoinf. 2022, 114, 103063. [Google Scholar] [CrossRef]
  13. Zeng, Z.; Lei, L.; Hou, S.; Ru, F.; Guan, X.; Zhang, B. A Regional Gap-Filling Method Based on Spatiotemporal Variogram Model of CO2 Columns. IEEE Trans. Geosci. Remote Sens. 2014, 52, 3594–3603. [Google Scholar] [CrossRef]
  14. Sheng, M.; Lei, L.; Zeng, Z.-C.; Rao, W.; Zhang, S. Detecting the Responses of CO2 Column Abundances to Anthropogenic Emissions from Satellite Observations of GOSAT and OCO-2. Remote Sens. 2021, 13, 3524. [Google Scholar] [CrossRef]
  15. Zhang, S.; Lei, L.; Sheng, M.; Song, H.; Li, L.; Guo, K.; Ma, C.; Liu, L.; Zeng, Z. Evaluating Anthropogenic CO2 Bottom-Up Emission Inventories Using Satellite Observations from GOSAT and OCO-2. Remote Sens. 2022, 14, 5024. [Google Scholar] [CrossRef]
  16. Uddin, M.S.; Czajkowski, K.P. Performance Assessment of Spatial Interpolation Methods for the Estimation of Atmospheric Carbon Dioxide in the Wider Geographic Extent. J. Geovis. Spat. Anal. 2022, 6, 10. [Google Scholar] [CrossRef]
  17. He, C.; Ji, M.; Li, T.; Liu, X.; Tang, D.; Zhang, S.; Luo, Y.; Grieneisen, M.L.; Zhou, Z.; Zhan, Y. Deriving Full-Coverage and Fine-Scale XCO2 Across China Based on OCO-2 Satellite Retrievals and CarbonTracker Output. Geophys. Res. Lett. 2022, 49, e2022GL098435. [Google Scholar] [CrossRef]
  18. Wu, C.; Ju, Y.; Yang, S.; Zhang, Z.; Chen, Y. Reconstructing Annual XCO2 at a 1 Km × 1 Km Spatial Resolution across China from 2012 to 2019 Based on a Spatial CatBoost Method. Environ. Res. 2023, 236, 116866. [Google Scholar] [CrossRef]
  19. Kallio, J.; Tervonen, J.; Räsänen, P.; Mäkynen, R.; Koivusaari, J.; Peltola, J. Forecasting Office Indoor CO2 Concentration Using Machine Learning with a One-Year Dataset. Build. Environ. 2021, 187, 107409. [Google Scholar] [CrossRef]
  20. Zhao, Z.; Xie, F.; Ren, T.; Zhao, C. Atmospheric CO2 Retrieval from Satellite Spectral Measurements by a Two-Step Machine Learning Approach. J. Quant. Spectrosc. Radiat. Transf. 2022, 278, 108006. [Google Scholar] [CrossRef]
  21. Finch, D.; Palmer, P.; Zhang, T. Automated Detection of Atmospheric NO2 Plumes from Satellite Data: A Tool to Help Infer Anthropogenic Combustion Emissions. Atmos. Meas. Tech. 2021, 15, 721–733. [Google Scholar] [CrossRef]
  22. Hakkarainen, J.; Ialongo, I.; Oda, T.; Szeląg, M.E.; O’Dell, C.W.; Eldering, A.; Crisp, D. Building a Bridge: Characterizing Major Anthropogenic Point Sources in the South African Highveld Region Using OCO-3 Carbon Dioxide Snapshot Area Maps and Sentinel-5P/TROPOMI Nitrogen Dioxide Columns. Environ. Res. Lett. 2023, 18, 035003. [Google Scholar] [CrossRef]
  23. Saw, G.K.; Dey, S.; Kaushal, H.; Lal, K. Tracking NO2 Emission from Thermal Power Plants in North India Using TROPOMI Data. Atmos. Environ. 2021, 259, 118514. [Google Scholar] [CrossRef]
  24. Fuentes Andrade, B.; Buchwitz, M.; Reuter, M.; Bovensmann, H.; Richter, A.; Boesch, H.; Burrows, J.P. A Method for Estimating Localized CO2 Emissions from Co-Located Satellite XCO2 and NO2 Images. Atmos. Meas. Tech. 2024, 17, 1145–1173. [Google Scholar] [CrossRef]
  25. Goldberg, D.L.; Lu, Z.; Oda, T.; Lamsal, L.N.; Liu, F.; Griffin, D.; McLinden, C.A.; Krotkov, N.A.; Duncan, B.N.; Streets, D.G. Exploiting OMI NO2 Satellite Observations to Infer Fossil-Fuel CO2 Emissions from U.S. Megacities. Sci. Total Environ. 2019, 695, 133805. [Google Scholar] [CrossRef]
  26. Konovalov, I.B.; Berezin, E.V.; Ciais, P.; Broquet, G.; Zhuravlev, R.V.; Janssens-Maenhout, G. Estimation of Fossil-Fuel CO2 Emissions Using Satellite Measurements of “Proxy” Species. Atmos. Chem. Phys. 2016, 16, 13509–13540. [Google Scholar] [CrossRef]
  27. Liu, F.; Duncan, B.N.; Krotkov, N.A.; Lamsal, L.N.; Beirle, S.; Griffin, D.; McLinden, C.A.; Goldberg, D.L.; Lu, Z. A Methodology to Constrain Carbon Dioxide Emissions from Coal-Fired Power Plants Using Satellite Observations of Co-Emitted Nitrogen Dioxide. Atmos. Chem. Phys. 2020, 20, 99–116. [Google Scholar] [CrossRef]
  28. He, Q.; Ye, T.; Wang, W.; Luo, M.; Song, Y.; Zhang, M. Spatiotemporally Continuous Estimates of Daily 1-Km PM2.5 Concentrations and Their Long-Term Exposure in China from 2000 to 2020. J. Environ. Manag. 2023, 342, 118145. [Google Scholar] [CrossRef]
  29. Park, H.; Jeong, S.; Park, H.; Labzovskii, L.D.; Bowman, K.W. An Assessment of Emission Characteristics of Northern Hemisphere Cities Using Spaceborne Observations of CO2, CO, and NO2. Remote Sens. Environ. 2021, 254, 112246. [Google Scholar] [CrossRef]
  30. Hakkarainen, J.; Ialongo, I.; Maksyutov, S.; Crisp, D. Analysis of Four Years of Global XCO2 Anomalies as Seen by Orbiting Carbon Observatory-2. Remote Sens. 2019, 11, 850. [Google Scholar] [CrossRef]
  31. Liu, D.; Lei, L.; Guo, L.; Zeng, Z.-C. A Cluster of CO2 Change Characteristics with GOSAT Observations for Viewing the Spatial Pattern of CO2 Emission and Absorption. Atmosphere 2015, 6, 1695–1713. [Google Scholar] [CrossRef]
  32. Wang, W.; Tian, Y.; Liu, C.; Sun, Y.; Liu, W.; Xie, P.; Liu, J.; Xu, J.; Morino, I.; Velazco, V.A.; et al. Investigating the Performance of a Greenhouse Gas Observatory in Hefei, China. Atmos. Meas. Tech. 2017, 10, 2627–2643. [Google Scholar] [CrossRef]
  33. Ciais, P.; Dolman, A.J.; Bombelli, A.; Duren, R.; Peregon, A.; Rayner, P.J.; Miller, C.; Gobron, N.; Kinderman, G.; Marland, G.; et al. Current Systematic Carbon-Cycle Observations and the Need for Implementing a Policy-Relevant Carbon Observing System. Biogeosciences 2014, 11, 3547–3602. [Google Scholar] [CrossRef]
  34. Yang, E.G.; Kort, E.A.; Ott, L.E.; Oda, T.; Lin, J.C. Using Space-Based CO 2 and NO 2 Observations to Estimate Urban CO2 Emissions. JGR Atmos. 2023, 128, e2022JD037736. [Google Scholar] [CrossRef]
  35. Ebi, K.L.; Anderson, C.L.; Hess, J.J.; Kim, S.-H.; Loladze, I.; Neumann, R.B.; Singh, D.; Ziska, L.; Wood, R. Nutritional Quality of Crops in a High CO2 World: An Agenda for Research and Technology Development. Environ. Res. Lett. 2021, 16, 064045. [Google Scholar] [CrossRef]
  36. Warren, J.M.; Jensen, A.M.; Ward, E.J.; Guha, A.; Childs, J.; Wullschleger, S.D.; Hanson, P.J. Divergent Species-specific Impacts of Whole Ecosystem Warming and Elevated CO2 on Vegetation Water Relations in an Ombrotrophic Peatland. Glob. Change Biol. 2021, 27, 1820–1835. [Google Scholar] [CrossRef]
  37. European Commission; Joint Research Centre. Fossil CO2 Emissions of All World Countries: 2018 Report; Publications Office: Luxembourg, 2018. [Google Scholar]
  38. He, Z.; Lei, L.; Welp, L.; Zeng, Z.-C.; Bie, N.; Yang, S.; Liu, L. Detection of Spatiotemporal Extreme Changes in Atmospheric CO2 Concentration Based on Satellite Observations. Remote Sens. 2018, 10, 839. [Google Scholar] [CrossRef]
  39. Chen, S.; Mihara, K.; Wen, J. Time Series Prediction of CO2, TVOC and HCHO Based on Machine Learning at Different Sampling Points. Build. Environ. 2018, 146, 238–246. [Google Scholar] [CrossRef]
  40. Kalra, S.; Lamba, R.; Sharma, M. Machine Learning Based Analysis for Relation between Global Temperature and Concentrations of Greenhouse Gases. J. Inf. Optim. Sci. 2020, 41, 73–84. [Google Scholar] [CrossRef]
  41. Siabi, Z.; Falahatkar, S.; Alavi, S.J. Spatial Distribution of XCO2 Using OCO-2 Data in Growing Seasons. J. Environ. Manag. 2019, 244, 110–118. [Google Scholar] [CrossRef]
  42. O’Dell, C.W.; Eldering, A.; Wennberg, P.O.; Crisp, D.; Gunson, M.R.; Fisher, B.; Frankenberg, C.; Kiel, M.; Lindqvist, H.; Mandrake, L.; et al. Improved Retrievals of Carbon Dioxide from Orbiting Carbon Observatory-2 with the Version 8 ACOS Algorithm. Atmos. Meas. Tech. 2018, 11, 6539–6576. [Google Scholar] [CrossRef]
  43. Massie, S.T.; Sebastian Schmidt, K.; Eldering, A.; Crisp, D. Observational Evidence of 3-D Cloud Effects in OCO-2 CO2 Retrievals. JGR Atmos. 2017, 122, 7064–7085. [Google Scholar] [CrossRef]
  44. Van Geffen, J.; Boersma, K.F.; Eskes, H.; Sneep, M.; Ter Linden, M.; Zara, M.; Veefkind, J.P. S5P TROPOMI NO2 Slant Column Retrieval: Method, Stability, Uncertainties and Comparisons with OMI. Atmos. Meas. Tech. 2020, 13, 1315–1335. [Google Scholar] [CrossRef]
  45. Fan, C.; Li, Z.; Li, Y.; Dong, J.; Van Der, A.R.; De Leeuw, G. Variability of NO2 Concentrations over China and Effect on Air Quality Derived from Satellite and Ground-Based Observations. Atmos. Chem. Phys. 2021, 21, 7723–7748. [Google Scholar] [CrossRef]
  46. Fioletov, V.; McLinden, C.A.; Griffin, D.; Krotkov, N.; Liu, F.; Eskes, H. Quantifying Urban, Industrial, and Background Changes in NO2 during the COVID-19 Lockdown Period Based on TROPOMI Satellite Observations. Atmos. Chem. Phys. 2022, 22, 4201–4236. [Google Scholar] [CrossRef]
  47. Veefkind, J.P.; Aben, I.; McMullan, K.; Förster, H.; De Vries, J.; Otter, G.; Claas, J.; Eskes, H.J.; De Haan, J.F.; Kleipool, Q.; et al. TROPOMI on the ESA Sentinel-5 Precursor: A GMES Mission for Global Observations of the Atmospheric Composition for Climate, Air Quality and Ozone Layer Applications. Remote Sens. Environ. 2012, 120, 70–83. [Google Scholar] [CrossRef]
  48. Guo, M.; Wang, X.; Li, J.; Wang, H.; Tani, H. Examining the Relationships between Land Cover and Greenhouse Gas Concentrations Using Remote-Sensing Data in East Asia. Int. J. Remote Sens. 2013, 34, 4281–4303. [Google Scholar] [CrossRef]
  49. Yang, W.; Zhao, Y.; Wang, Q.; Guan, B. Climate, CO2, and Anthropogenic Drivers of Accelerated Vegetation Greening in the Haihe River Basin. Remote Sens. 2022, 14, 268. [Google Scholar] [CrossRef]
  50. Chen, X.; He, Q.; Ye, T.; Liang, Y.; Li, Y. Decoding Spatiotemporal Dynamics in Atmospheric CO2 in Chinese Cities: Insights from Satellite Remote Sensing and Geographically and Temporally Weighted Regression Analysis. Sci. Total Environ. 2024, 908, 167917. [Google Scholar] [CrossRef]
  51. Buchwitz, M.; Reuter, M.; Schneising, O.; Noël, S.; Gier, B.; Bovensmann, H.; Burrows, J.P.; Boesch, H.; Anand, J.; Parker, R.J.; et al. Computation and Analysis of Atmospheric Carbon Dioxide Annual Mean Growth Rates from Satellite Observations during 2003–2016. Atmos. Chem. Phys. 2018, 18, 17355–17370. [Google Scholar] [CrossRef]
  52. Sun, Y.; Liu, C.; Zhang, L.; Palm, M.; Notholt, J.; Yin, H.; Vigouroux, C.; Lutsch, E.; Wang, W.; Shan, C.; et al. Fourier Transform Infrared Time Series of Tropospheric HCN in Eastern China: Seasonality, Interannual Variability, and Source Attribution. Atmos. Chem. Phys. 2020, 20, 5437–5456. [Google Scholar] [CrossRef]
  53. Hersbach, H.; Bell, B.; Berrisford, P.; Hirahara, S.; Horányi, A.; Muñoz-Sabater, J.; Nicolas, J.; Peubey, C.; Radu, R.; Schepers, D.; et al. The ERA5 Global Reanalysis. Quart. J. R. Meteorol. Soc. 2020, 146, 1999–2049. [Google Scholar] [CrossRef]
  54. Jiang, Z.; Huete, A.R.; Chen, J.; Chen, Y.; Li, J.; Yan, G.; Zhang, X. Analysis of NDVI and Scaled Difference Vegetation Index Retrievals of Vegetation Fraction. Remote Sens. Environ. 2006, 101, 366–378. [Google Scholar] [CrossRef]
  55. Chatterjee, A.; Payne, V.; Eldering, A.; Rosenberg, R.; Kiel, M.; Fisher, B.; Nelson, R.; Dang, L.; Rodrigues, G.K.; O’Dell, C.; et al. Orbiting Carbon Observatory-3 (OCO-3) Data Quality Statement: Level 2 Forward and Retrospective Processing Data Release 10 (V10 and V10r), V10.4 Lite Files; California Institute of Technology: Pasadena, CA, USA, 2022. [Google Scholar]
  56. Keely, W.R.; Mauceri, S.; Crowell, S.; O’Dell, C.W. A Nonlinear Data-Driven Approach to Bias Correction of XCO 2 for NASA’s OCO-2 ACOS Version 10. Atmos. Meas. Tech. 2023, 16, 5725–5748. [Google Scholar] [CrossRef]
  57. Zhang, J. China Greenhouse Gas Bulletin 2022 post. China Meteorological News, 6 December 2023. [Google Scholar]
Figure 1. Framework for predicting XCO2 based on ML, involving reconstruction XCO2 data and analysis of the NO2 field to constrain XCO2 predictions.
Figure 1. Framework for predicting XCO2 based on ML, involving reconstruction XCO2 data and analysis of the NO2 field to constrain XCO2 predictions.
Remotesensing 16 02456 g001
Figure 2. Validations of the model predicting XCO2 through (a) cross-validation and (b) comparison with T-XCO2.
Figure 2. Validations of the model predicting XCO2 through (a) cross-validation and (b) comparison with T-XCO2.
Remotesensing 16 02456 g002
Figure 3. Spatial distribution of the mean XCO2 values in the study area. (a) Predicted XCO2 (Cm-XCO2) and (b) fp-XCO2 for each 0.1° grid.
Figure 3. Spatial distribution of the mean XCO2 values in the study area. (a) Predicted XCO2 (Cm-XCO2) and (b) fp-XCO2 for each 0.1° grid.
Remotesensing 16 02456 g003
Figure 4. Effects of the predictive model for the original satellite-observed XCO2 (fp-XCO2) that was re-calculated using this model. (a) Comparison of the co-located fp-XCO2 and Cm-XCO2 based on the monthly averages of the XCO2 retrievals within a 0.1° grid from 2019 to 2022. (b) Histograms of the co-located fp-XCO2 and Cm-XCO2. (c) Mean difference between the co-located Cm-XCO2 and fp-XCO2 between 2019 and 2022 in the study area. (d) Mean posterior uncertainty in XCO2 from the L2 product algorithm.
Figure 4. Effects of the predictive model for the original satellite-observed XCO2 (fp-XCO2) that was re-calculated using this model. (a) Comparison of the co-located fp-XCO2 and Cm-XCO2 based on the monthly averages of the XCO2 retrievals within a 0.1° grid from 2019 to 2022. (b) Histograms of the co-located fp-XCO2 and Cm-XCO2. (c) Mean difference between the co-located Cm-XCO2 and fp-XCO2 between 2019 and 2022 in the study area. (d) Mean posterior uncertainty in XCO2 from the L2 product algorithm.
Remotesensing 16 02456 g004
Figure 5. Contribution of the predictor variables in modeling XCO2 using the CatB algorithm.
Figure 5. Contribution of the predictor variables in modeling XCO2 using the CatB algorithm.
Remotesensing 16 02456 g005
Figure 6. The relationship between NO2 and anthropogenic CO2 emissions from EDGAR from 2019 to 2022. (a) Annual mean atmospheric NO2 concentration. (b) Annual mean anthropogenic CO2 emissions (2019–2022). (c) NO2 and XCO2 response to anthropogenic CO2 emissions from EDGAR calculated based on the averaged values of the grids within each clustered area derived from the spatio-temporal clustering analysis of NO2 data.
Figure 6. The relationship between NO2 and anthropogenic CO2 emissions from EDGAR from 2019 to 2022. (a) Annual mean atmospheric NO2 concentration. (b) Annual mean anthropogenic CO2 emissions (2019–2022). (c) NO2 and XCO2 response to anthropogenic CO2 emissions from EDGAR calculated based on the averaged values of the grids within each clustered area derived from the spatio-temporal clustering analysis of NO2 data.
Remotesensing 16 02456 g006
Figure 7. The relationship between NO2 and XCO2 calculated based on the averaged values of the grids within each clustered area derived from the spatio-temporal clustering analysis of NO2 data.
Figure 7. The relationship between NO2 and XCO2 calculated based on the averaged values of the grids within each clustered area derived from the spatio-temporal clustering analysis of NO2 data.
Remotesensing 16 02456 g007
Figure 8. Analysis of NO2 and XCO2 in regions of interest (ROIs) with (a) ROI locations where the background map shows the clustering results of NO2 data from 2019 to 2022 and the legend presents the 14 classes and averaged values of NO2 from 2019 to 2022 for each class to the left and right of symbol, (b) yearly average values of ΔNO2 and ΔXCO2 (y-axis) for each ROI (x-axis) in the winter (December–February) for the four years of 2019–2022, and (c) the relationship between NO2 and XCO2, which is calculated by using 2019 yearly averages as contrasting values for each ROI in the three years 2020, 2021, and 2022, respectively.
Figure 8. Analysis of NO2 and XCO2 in regions of interest (ROIs) with (a) ROI locations where the background map shows the clustering results of NO2 data from 2019 to 2022 and the legend presents the 14 classes and averaged values of NO2 from 2019 to 2022 for each class to the left and right of symbol, (b) yearly average values of ΔNO2 and ΔXCO2 (y-axis) for each ROI (x-axis) in the winter (December–February) for the four years of 2019–2022, and (c) the relationship between NO2 and XCO2, which is calculated by using 2019 yearly averages as contrasting values for each ROI in the three years 2020, 2021, and 2022, respectively.
Remotesensing 16 02456 g008
Figure 9. An example of differences between the monthly XCO2 and NO2 values from 2020 to 2022 and the same month in 2019 for R1-Wuhan, R2-Shngahia, R3-YRD, and 4-BTH, as well as the overall study area.
Figure 9. An example of differences between the monthly XCO2 and NO2 values from 2020 to 2022 and the same month in 2019 for R1-Wuhan, R2-Shngahia, R3-YRD, and 4-BTH, as well as the overall study area.
Remotesensing 16 02456 g009
Figure 10. The means and the differences of fp-XCO2 from 2019 to 2022 derived from (a,d) LGB, (b,e) MLP, and (c,f) LSTM.
Figure 10. The means and the differences of fp-XCO2 from 2019 to 2022 derived from (a,d) LGB, (b,e) MLP, and (c,f) LSTM.
Remotesensing 16 02456 g010
Figure 11. Machine learning model SHAP value swarm graphs, (a) LGB, (b) MLP, and (c) LSTM.
Figure 11. Machine learning model SHAP value swarm graphs, (a) LGB, (b) MLP, and (c) LSTM.
Remotesensing 16 02456 g011
Table 1. Summary table of the multisource data used.
Table 1. Summary table of the multisource data used.
AcronymParameterSourceResolutionProduct
SpaceTime
fp-XCO2XCO2 retrievalsOCO-22.25 km × 1.29 km16 daysOCO-2
_L2_Lite_FP_11r
OCO-32.25 km × 1.29 km16 daysOCO-3
_L2_Lite_FP_10.4r
NO2Atmospheric NO2 columnTROPOMI-S5P0.01°MonthlySentinel-5P OFFL NO2: Offline Nitrogen Dioxide
NDVINormalized-difference vegetation indexMODIS0.05°MonthlyMOD13C2
D2M2 m dewpoint temperatureERA5—
fifth-generation ECMWF atmospheric reanalysis
0.1°MonthlyComplete ERA5 global atmospheric reanalysis
T2M2 m temperature
U1010 m U-wind component
V1010 m V-wind component
MXCO2Mapping XCO2Mapped geostatistical method using XCO2 retrievals0.5°MonthlyGlobal land 0.5° mapping XCO2 dataset using satellite observations of GOSAT, OCO-2, and OCO-3 from 2009 to 2022
T-XCO2Ground-based XCO2 dataTCCONPoint-TCCON data from Hefei (PRC)
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

Guo, K.; Lei, L.; Sheng, M.; Ji, Z.; Song, H. Refining Spatial and Temporal XCO2 Characteristics Observed by Orbiting Carbon Observatory-2 and Orbiting Carbon Observatory-3 Using Sentinel-5P Tropospheric Monitoring Instrument NO2 Observations in China. Remote Sens. 2024, 16, 2456. https://doi.org/10.3390/rs16132456

AMA Style

Guo K, Lei L, Sheng M, Ji Z, Song H. Refining Spatial and Temporal XCO2 Characteristics Observed by Orbiting Carbon Observatory-2 and Orbiting Carbon Observatory-3 Using Sentinel-5P Tropospheric Monitoring Instrument NO2 Observations in China. Remote Sensing. 2024; 16(13):2456. https://doi.org/10.3390/rs16132456

Chicago/Turabian Style

Guo, Kaiyuan, Liping Lei, Mengya Sheng, Zhanghui Ji, and Hao Song. 2024. "Refining Spatial and Temporal XCO2 Characteristics Observed by Orbiting Carbon Observatory-2 and Orbiting Carbon Observatory-3 Using Sentinel-5P Tropospheric Monitoring Instrument NO2 Observations in China" Remote Sensing 16, no. 13: 2456. https://doi.org/10.3390/rs16132456

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