Next Article in Journal
Choice of Solar Spectral Irradiance Model for Current and Future Remote Sensing Satellite Missions
Previous Article in Journal
Removing Moving Objects without Registration from 3D LiDAR Data Using Range Flow Coupled with IMU Measurements
Previous Article in Special Issue
Evaluation of the Accuracy of Spectral Calibration Light Source on Spectral Radiance Acquired by the Greenhouse-Gases Absorption Spectrometer-2 (GAS-2)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Spatial and Temporal Variations of Atmospheric CH4 in Monsoon Asia Detected by Satellite Observations of GOSAT and TROPOMI

1
School of Earth Sciences and Resources, China University of Geosciences Beijing, Beijing 100083, China
2
School of Earth and Space Sciences, Peking University, Beijing 100871, China
3
China Highway Engineering Consultants Corporation, Beijing 100089, China
4
Key Laboratory of Digital Earth Science, Aerospace Information Research Institute, Chinese Academy of Sciences, Beijing 100094, China
5
Big Data for Sustainable Development Goals, Beijing 100094, China
*
Author to whom correspondence should be addressed.
Remote Sens. 2023, 15(13), 3389; https://doi.org/10.3390/rs15133389
Submission received: 11 April 2023 / Revised: 2 June 2023 / Accepted: 7 June 2023 / Published: 3 July 2023
(This article belongs to the Special Issue Remote Sensing of Greenhouse Gas Emissions)

Abstract

:
Space-based measurements, such as the Greenhouse gases Observing SATellite (GOSAT) and the TROPOspheric Monitoring Instrument (TROPOMI) aboard the Sentinel-5 Precursor satellite, provide global observations of the column-averaged CH4 concentration (XCH4). Due to the irregular observations and data gaps in the retrievals, studies on the spatial and temporal variations of regional atmospheric CH4 concentrations are limited. In this paper, we mapped XCH4 data over monsoon Asia using GOSAT and TROPOMI observations from April 2009 to December 2021 and analyzed the spatial and temporal pattern of atmospheric CH4 variations and emissions. The results show that atmospheric CH4 concentrations over monsoon Asia have long-term increases with an annual growth rate of roughly 8.4 ppb. The spatial and temporal trends of XCH4 data are significantly correlated with anthropogenic CH4 emissions from the bottom-up emission inventory of EDGAR. The spatial pattern of gridded XCH4 temporal variations in China presents a basically consistent distribution with the Heihe–Tengchong Line, which is mainly related to the difference in anthropogenic emissions in the eastern and western areas. Using the mapping of XCH4 data from 2019 to 2021, this study further revealed the response of atmospheric CH4 concentrations to anthropogenic emissions in different urban agglomerations. For the urban agglomerations, the triangle of Central China (TCC), the Chengdu–Chongqing City Group (CCG), and the Yangtze River Delta (YRD) show higher CH4 concentrations and emissions than the Beijing–Tianjin–Hebei region and nearby areas (BTH). The results reveal the spatial and temporal distribution of CH4 concentrations and quantify the differences between urban agglomerations, which will support further studies on the drivers of methane emissions.

Graphical Abstract

1. Introduction

Atmospheric CH4 is the most important contributor to the anthropogenically enhanced greenhouse warming effect after CO2 [1]. According to the IPCC Sixth Assessment Report (IPCC AR6), the global warming potential of CH4 is 28 times greater than that of CO2 in 100 years [2,3]. The oxidation of methane eventually leads to CO2. As CH4 decays, it also leads to the formation of tropospheric O3 and stratospheric H2O and changes in the concentrations of tropospheric hydroxyl, resulting in indirect radiative forcing [4]. The budget of CH4 consists of a diverse mix of sources and sinks, with many sources overlapping spatially, so it is difficult to quantify the emissions [5]. Anthropogenic source emission types account for roughly 60% of atmospheric CH4 emissions, such as agricultural activities, energy activities, and waste management [6,7,8]. Natural source emission types account for about 40%, including wetlands, oceans, termites, and clathrates [9]. In the background of tackling climate change and achieving a dual carbon target (aiming to have carbon dioxide emissions peak before 2030 and achieve carbon neutrality before 2060), monitoring regional CH4 concentrations and their changes is important to help us better understand the driving factors of CH4 variations and estimate CH4 emissions.
Ground-based observations have become an effective method for revealing the relationship between greenhouse gases and anthropogenic emissions [10]. The long-term, high-quality surface observations of CH4 and δ13C (CH4) abundances by the WMO GAW community show that the global average atmospheric CH4 has been increasing at an accelerating rate since 2007. Among them, the annual increases in 2020 and 2021 were the largest since the systematic record began in 1983 [11]. The main reason for the continuous increase in atmospheric CH4 concentrations is the increase in anthropogenic emissions. Previous studies have suggested that this is likely to be caused by anthropogenic emissions from agricultural and fossil fuel sources in the tropics and the northern mid-latitudes and the decline in the atmospheric sink [3,12]. Although the atmospheric CH4 concentration observed on the ground is highly accurate, it is difficult to identify and quantify the spatiotemporal changes in methane sources and sinks on a regional scale with sparsely distributed station observations.
Greenhouse gases Observing Satellite (GOSAT) and Copernicus Sentinel-5 Precursor (Sentinel-5P) observations provide important data sources for studying the spatiotemporal changes in the global atmospheric CH4 concentration. GOSAT was successfully launched in January 2009 and is the first satellite dedicated to greenhouse gas observation. It was jointly developed by the Japan Aerospace Exploration Agency (JAXA), the National Institute of Environment (NIES), and the Ministry of Environment (MOE). Its main purpose is to obtain global atmospheric CO2 and CH4 concentration distribution information from space and estimate carbon sources and sinks on a continental scale, providing an important basis for formulating carbon emission reduction policies. The satellite observations of GOSAT cover the world within 3 days and have a spatial resolution of roughly 10.5 km at their nadir. The Copernicus Sentinel-5 Precursor mission consists of one satellite carrying the TROPOspheric Monitoring Instrument (TROPOMI) instrument launched by ESA on 13 October 2017. It can effectively observe trace gas components (such as CH4, NO2, CO, etc.) around the world on an urban scale. The width swath of TROPOM is approximately 2600 km on the ground, and the resolution is up to 7 km × 3.5 km, which greatly improves the accuracy of greenhouse gas observations with full daily global surface coverage. Previous studies have shown that these satellite observation data have the application potential of detecting emission sources and estimating anthropogenic emissions from regional or point sources [13,14,15]. However, due to the influence of cloud cover and sensor observation mode, the CH4 concentration data observed by satellite remote sensing are irregularly distributed in time and space.
In order to obtain effective CH4 concentration data with as much global spatiotemporal continuous distribution as possible, scholars use the kriging statistical method to interpolate the satellite observation data. Tomosada et al. [16] used the XCO2 data product retrieved by GOSAT for kriging interpolation, calculated the variation function by separating the ocean and land, and considered the anisotropy of CO2 concentration distribution to establish a semi-variation model. Bruhwiller et al. [6] used the general kriging method with a moving window to output the spatiotemporal continuous distribution data of global atmospheric CH4 with a high temporal resolution. The observed mean mole fraction of CH4 in the latitude zone was simulated, and the interannual variation of CH4 emissions in the high latitude region (53°–90°N) was captured. Liu et al. [17] obtained the daily mean value of XCO2 and XCH4 in East Asia by using the L2 data products of XCO2 and XCH4 retrieved from the single-year GOSAT satellite observation, with the spatial kriging interpolation method. The overall result is consistent with the L3 product data released by the GOSAT satellite standard. With the accumulation of geographic spatiotemporal data and the development of the theory of spatiotemporal variation function, geostatistic methods have also been extended from space to the use of spatiotemporal interpolation. Liu and Li et al. [18,19] carried out global or regional XCH4 mapping analysis using single satellite observation data based on spatiotemporal and geostatistical methods, and the results showed that the spatiotemporal distribution of regional CH4 concentrations was correlated with the anthropogenic emission inventory, especially in eastern China, Bangladesh, the Indo-China Peninsula, and nearby regions. However, the observation time series and effective data points of a single satellite are limited in these studies. Using the fusion mapping results of multiple satellites can more effectively reveal the long-term dynamic changes in CH4 concentrations.
The CH4 concentration in the Asian monsoon region has continued to increase in recent years. However, due to the diversity and complex distribution of CH4 emission sources in the Asian region, with a variety of climate types and land cover types, the response mechanism between the CH4 emission and the space–time distribution of CH4 concentrations in the region is still not clear [2], in which case, further analysis is needed. This study generates the regional long-term XCH4 spatiotemporal continuous data set (Mapping-XCH4) based on the spatiotemporal statistical method, which is able to utilize both time and space information simultaneously, using the XCH4 data observed by GOSAT and TROPOMI from 2009 to 2021. The spatiotemporal variation trend of XCH4 in the Asian monsoon region is studied on different spatial scales. This paper aims to use XCH4 spatiotemporal continuous data to detect and analyze the XCH4 spatiotemporal changes in the Asian monsoon region at different spatial scales, reveal the application ability of satellite mapping data in the study of XCH4 spatiotemporal characteristics, and provide effective support data for further research on the driving factors of atmospheric CH4 concentration changes and carbon sources and sinks in the Asian monsoon region.

2. Materials and Methods

2.1. Materials

2.1.1. Satellite XCH4 Observations

To analyze the spatial and temporal distribution characteristics of atmospheric CH4 concentrations in the Asian monsoon region, we collected satellite-observed XCH4 data from GOSAT and S5P spanning from April 2009 to December 2021. Table 1 gives a summary of the XCH4 data products.
The XCH4 data from GOSAT observations (GOSAT-XCH4) are the bias-corrected FTS SWIR Level 2 CH4 products (V02.95/V02.96) distributed by the National Institute for Environmental Studies (NIES), which are provided from the GOSAT Data Archive Service (GDAS) as monthly archive data (https://data2.gosat.nies.go.jp/GosatDataArchiveService/usr/download, (accessed on 1 March 2022)). GOSAT Project. Available online: www.gosat.nies.go.jp (accessed on 1 June 2020). According to the bias-correction processing results referred to in [20], the differences between the corrected SWIR L2 CH4 product and the TCCON data from 2009 to 2018 are −0.140 ± 11.601 ppb for Gain H and 0.00 ± 18.428 ppb for Gain M over land.
The XCH4 data observed by TROPOMI (TROPOMI-XCH4) are from the Level 2 Offline (OFFL) timeliness data products that are available on the Copernicus S5p Open Access Hub (https://s5phub.copernicus.eu/dhus/#/home (accessed on 5 March 2022)). The detailed descriptions of CH4 OFFL data versions, the respective dates, and the relevant improvements can be found in the Product Readme File (PRF) (https://sentinels.copernicus.eu/documents/247904/3541451/Sentinel-5P-Methane-Product-Readme-File.pdf/d7214038-25a9-416f-8deb-d5d6c766f92e?t=1658386700684 (accessed on 7 March 2022)). The XCH4 retrieval is in good overall agreement with correlative ground-based measurements from TCCON and the Network for the Detection of Atmospheric Composition Change (NDACC) [21,22]. The systematic differences of the bias-corrected XCH4 data with respect to TCCON data and NDACC data are, on average, −0.26 ± 0.56% and 0.57 ± 0.83%, respectively [23].

2.1.2. Ground-Based XCH4 Data

Ground-based XCH4 measurements of the Total Carbon Column Observing Network (TCCON) are used to validate XCH4 data from Mapping-XCH4 and individual satellite observations of GOSAT and S5P. TCCON is a network recording direct solar spectra in the NIR/SWIR spectral region using ground-based Fourier-transform spectrometers (FTSs). Accurate and precise column-averaged abundances of CH4 retrieved from these high spectral (0.02 cm−1) and temporal (~90 s) resolution spectra are the primary data for validating satellite XCH4 data products.
There are four TCCON sites in the monsoon region of Asia, including Saga (Shiomi et al. [24]), Rikubetsu (Morino et al. [25]), Tsukuba (Morino et al. [26]), Xianghe (Yang et al. [27]), and Hefei (Liu et al. [28]). We used the GGG2020 version of the TCCON data products available from the TCCON data archive (https://tccondata.org/, accessed on 15 July 2022). In order to match the temporal scale of the Mapping-XCH4 data, TCCON XCH4 observations within 13:30 ± 2 h of the local standard time are selected to calculate the 3-day mean values.

2.1.3. CH4 Emission Data from EDGAR

CH4 emissions from the Emissions Database for Global Atmospheric Research (EDGAR) are used to compare with spatial XCH4 patterns derived from satellite observations. EDGAR is a global greenhouse gas inventory developed by the European Commission’s Joint Research Centre (JRC) based on the bottom-up emissions calculation methodology using international activity data [1,29,30]. The global emissions are spatially allocated on 0.1° × 0.1° resolution grids using proxy data, such as point sources, population density, traffic networks, and nighttime lights [29,30,31]. The new version V7.0 of the EDGAR emission inventory provides global CH4 estimates from 1970 to 2021, which are disaggregated to IPCC-relevant source-sector levels [32]. We used annual grid maps provided in netCDF format in which CH4 emissions are expressed in kg substance/m2/s.

2.2. Methods

2.2.1. The XCH4 Mapping Method Using GOSAT and TROPOMI Observations

We used the XCH4 data observed from GOSAT and S5P to produce a spatiotemporally continuous mapping XCH4 dataset (Mapping-XCH4) in the monsoon region of Asia from April 2009 to December 2021. The dataset covers the Asian monsoon region that ranges from 10°N to 55°N and 61°E to 146°E, with a 1° grid interval in space and a 3-day interval in time. Figure 1 shows the flowchart of generation, which includes the steps of data preprocessing, spatiotemporal correlation analysis, and the mapping of gridded XCH4 data.
(1)
Data preprocessing
To obtain valid XCH4 data over land with good-quality observations, the XCH4 data products were filtered using the criteria listed in Table 1. According to the geographic range of the study area, only XCH4 data within the buffer zone from 5°N to 60°N and 55°E to 150°E were used in this paper.
As shown in Figure 2, these XCH4 observations are irregularly distributed in space and time, with many gaps caused by impacts from geophysical factors such as aerosols and clouds. The spatial and temporal scales of GOSAT and TROPOMI observations are different, which results in the valid observations of GOSAT being much smaller than those of TROPOMI. Therefore, we converted the spatiotemporal scales of individual satellite XCH4 data by calculating the average value XCH4 within 3 days and 10 km. The equation is as follows:
X C H 4   u n i = i = 1 n X C H 4   o r i N
where X C H 4   o r i represents the XCH4 data with original spatiotemporal units (10.5 km/3 days for GOSAT, 7 km/1 day for TROPOMI), and N is the amount of X C H 4   o r i data within 3 days and a radius of 5 km. X C H 4   u n i represents the XCH4 data with the same space–time units (10 km/3 days).
This processing of converting spatiotemporal scales addresses the problem that GOSAT and TROPOMI observations have different sampling times at the same location while reducing the discrepancy between the XCH4 data. Figure 3 is the scatter plot of GOSAT-XCH4 and TROPOMI-XCH4 during the overlap period of 2019 to 2021. Comparing the XCH4 data at the original spatiotemporal scales, the overall difference between the GOSAT and S5P data is −0.63 ± 18.13 ppb, and the correlation coefficient (R) is 0.86. After converting the spatial and temporal scales, the data difference changes to a value of −0.23 ± 16.06 ppb, and the correlation is slightly improved (R = 0.87).
(2)
Spatiotemporal correlation analysis
The spatiotemporal correlation analysis aims to separate spatiotemporal deterministic trends and stochastic residual components from regional variables of XCH4, and further characterize the spatial and temporal correlations of XCH4 data in the study area [13,14,19]. In this paper, we applied the curve fitting method to long-term XCH4 data in each 2° latitude band, which referred to Sheng et al. [13]. Figure 4a is the spatiotemporal distribution of XCH4 deterministic trend, which shows annual increases, seasonal changes in time, and latitudinal gradient in space. The residual components of XCH4 are standard Normal distribution with a distribution range of ± 70 ppb.
Subsequently, we calculated the spatiotemporal variograms using 50 km and 3 days as the spatial and temporal steps, respectively. Figure 4b shows the spatiotemporal variograms model. In order to ensure the XCH4 data within the large-scale study area would be temporally and spatially autocorrelated, the spatial and temporal ranges that the model flattens out were set to 2500 km and 150 days. In this model, the nugget/sill ratio of about 0.11 indicates a strong spatial and temporal correlation of XCH4 data in the monsoon region in Asia. Compared to the previous studies, it outperforms the correlation calculated using only GOSAT XCH4 data in the Eurasia region [19].
(3)
The mapping of gridded XCH4 data
Based on the XCH4 residual values and the variogram model, we mapped the gridded data using the spatiotemporal kriging method, which has been widely used for global and regional greenhouse gas (CO2 and CH4) interpolation [13,14,18,19,33,34]. The mapping method uses the known XCH4 observations within a moving cylinder kriging neighborhood to predict the XCH4 values at the center of each 1° grid. In this paper, the initial ranges of the cylinder were set to 200 km in space and 15 time units (45 days) in time. If the number of available data was less than 20, the spatiotemporal ranges increased by 10 km and 1 time unit, respectively. In the mapping results, the maximum ranges were basically less than 40 time units and 450 km for all grids.

2.2.2. Local Indicators of Spatial Association

In this paper, we used the local indicators of spatial association (LISA) to conduct a spatial analysis of the CH4 concentration distribution in the study area to identify the spatial patterns of high-concentration and low-concentration CH4 areas. The local Moran’s I coefficient was utilized to evaluate the significance of the statistical results of each grid, and a proportional relationship was established between the local statistical data and the corresponding global statistical data [35]. The calculation method is shown in Equation (2):
I i = y i y 1 n y i y 2 j i n w i j y j y
where y i and y j are the XCH4 values of grid i and grid j , w i j is the spatial weight value of XCH4 between different grids calculated by the inverse distance squared method, n is the total number of all grids in the study area, and I represents the local Moran index of the i th grid. I i > 0 indicates a positive spatial correlation; the larger value shows the significant spatial correlation. I i < 0 indicates a negative spatial correlation; the smaller value shows the larger spatial disparity [36].

2.2.3. Clustering Spatial Pattern of XCH4 Temporal Changes

We studied the spatial aggregation characteristics of XCH4 temporal changes by clustering the XCH4 time series curves of each grid in the study area [13]. Particle swarm optimization (PSO) is used to optimize the clustering center of KNN (K-Nearest Neighbor) to obtain the optimal clustering center (30 categories) [37,38]. Within the same clustering category, the temporal variation characteristics of CH4 for each grid are the same or similar.

3. Results

3.1. Spatiotemporal Characteristic of XCH4 Variations in the Asian Monsoon Region from 2009 to 2021

The spatial and temporal variations of the Mapping-XCH4 data in the Asian monsoon region are shown in Figure 5 and Figure 6. The temporal characteristics of XCH4 variations in the study areas are extracted using the curve fitting method [13]. The long-term trend of XCH4 from 2009 to 2021 increases from 1785.63 ppb to 1897.28 ppb with an annual growth rate of about 8.4 ppb. There is an obvious latitudinal gradient in the XCH4, and the growth trends are similar in each latitudinal zone. Generally, the CH4 concentrations reach the highest value in August–November, and the lowest value in April–June each year. In Figure 6, the seasonal cycle peak varies with the latitude, because of the wide range of climate zones in the regions in Asia. Both the growth rates and the seasonal fluctuations of XCH4 from 2019 to 2021 are higher than those observed from 2009 to 2019, which may be related to different data sources used for mapping. Compared with the individual observations of GOSAT and TROPOMI, the mapping data fill the gap area in the satellite observations, especially for the period of GOSAT observations. The continuous XCH4 data in space and time reveal the regional spatiotemporal variation characteristics more comprehensively.
Figure 7 shows the spatial distribution of multi-year average XCH4 from 2010 to 2021. The low values are mainly distributed in China’s Qinghai–Tibet Plateau and the northern regions at latitudes above 40°N. The elevated CH4 concentrations are mostly located in the North China Plain, southern China, South Asia, and Southeast Asia. These regions are dotted with wetlands and rice paddy fields, which are primary sources of atmospheric CH4. Figure 8 shows that there are similar spatial distributions of the seasonally averaged XCH4 over the period from 2009 to 2021. CH4 concentrations in autumn and summer are generally higher than those in spring and winter, because of the stronger emissions from rice paddies and wetlands in the warmer and wetter seasons. The spatial pattern of XCH4 is consistent with the mapping result using GOSAT data from 2009 to 2014, as described in Liu et al. [18]. However, the long-term XCH4 data can reveal the more obvious seasonal variations of XCH4, especially differentiating CH4 concentrations between summer and autumn.

3.2. Spatial Pattern of XCH4 Variations Corresponding to the Surface Emissions

Figure 9 is the spatial distribution of local Moran’s I coefficients calculated based on multi-year averages from 2010 to 2021, which represents the spatial associations of XCH4 in the monsoon region of Asia. The dashed areas have a significant p-value of less than 0.05. The red and blue boxes correspond to the regions with low and high CH4 concentrations in Figure 7, respectively. The results indicate that the gridded XCH4 data for these regions have a significant spatial correlation. Compared to the nearby areas, the Sichuan Basin in China, northern India, and Bangladesh present extremely high XCH4 values due to high anthropogenic and natural emissions, such as oil deposits, paddy fields, livestock, and topography [39,40,41,42,43,44,45].
The clustering results of the gridded XCH4 time series are shown in Figure 10. The gridded XCH4 changes in the same cluster have a similar temporal trend that consists of long-term growth trends and seasonal cycles. The clusters present intense spatial aggregation; the regions with high CH4 concentrations are divided into clusters 4, 7, 10, 15, and 27, and those with low CH4 concentrations are divided into clusters 21, 24, 25, and 29. In the clusters of China, we noted that a clear east–west boundary is roughly consistent with the Heihe–Tengchong Line, which reflects two roughly equal parts with contrasting population densities. The XCH4 in the two parts has climbed steadily at a rate of ~8 ppb every year, but the concentrations in the eastern half are about 2% (25~45 ppb) higher than those in the west. The results indicate that the spatial pattern of CH4 in China is significantly affected by anthropogenic and natural emissions.
Compared to the anthropogenic emissions from EDGAR in Figure 11, the spatial pattern of high CH4 concentrations agrees well with the high-emission regions. At the same time, the CH4 emissions in China also showed obvious differences between the east and west parts. We calculated the spatial correlation between EDGAR emissions and Mapping-XCH4 based on the clusters to analyze the response of atmospheric CH4 concentrations to CH4 emissions. In Figure 11b, the XCH4 data of each cluster region are subtracted from the annual average of China. There is an exponential linear relationship between XCH4 and anthropogenic emissions in spatial distribution, and the correlation coefficients are greater than 0.7 for each year from 2010 to 2021. However, influenced by atmospheric transport, high anthropogenic emissions do not imply high CH4 concentrations. The clusters of 4, 5, 10, and 27 show emission values above the fitting trend.
For the temporal correlation, we compared the annual time series of XCH4 and anthropogenic emissions in these clusters. The regional CH4 concentrations increase linearly with anthropogenic emissions (R > 0.8) except for clusters 3, 18, and 29, which have low emissions and high topography. This is because changes in CH4 concentrations depend not only on surface emissions and sinks from chemical reactions but also on atmospheric transport caused by global and local circulations [44].

3.3. Regional Characteristics of XCH4 Variations in Urban Agglomerations of China

To investigate the regional characteristics of spatiotemporal XCH4 variations, we selected four urban agglomerations with high CH4 concentrations in China as the study areas (Figure 12), including the Beijing–Tianjin–Hebei region and nearby areas (BTH), the Yangtze River Delta (YRD), the triangle of Central China (TCC), and the Chengdu–Chongqing City Group (CCG).
Figure 13 presents the monthly time series of the regional-averaged XCH4 values calculated by the Mapping-XCH4 data (red line) and compared with regional fitting values (Fitting-XCH4) by the curve fitting method using raw satellite XCH4 observations (gray line). In urban agglomerations with a large number of available observations (BTH, YRD, and TCC), the temporal variations of the monthly values of Mapping-XCH4 and Fitting-XCH4 are generally consistent. The large differences only occurred in the summer months, with more missing data during the GOSAT observation period. In CCG, where there are relatively few satellite observations, the Mapping-XCH4 data also reflect the overall trend of XCH4 changes. Comparing the monthly time series of the Mapping-XCH4 and Fitting-XCH4 data, the large deviations are during the summer months in YRD and TCC and around 2012 in CCG. This is because the curve fitting method is susceptible to extreme values, which result in the Fitting-XCH4 data during the GOSAT observation period becoming more prone to outliers that are influenced by the spatial-temporal heterogeneity of the data. Therefore, the Mapping-XCH4 data can be applied to provide effective information on regional XCH4 changes.
We analyzed the variations in atmospheric CH4 concentrations in different urban agglomerations by comparing the regional characteristics of XCH4 and anthropogenic emissions from the EDGAR inventory (Figure 12). The CH4 concentrations in YRD, TCC, and CCG are higher than those in BTH, which are about 6.74 ± 2.17 ppm, 7.80 ± 3.71 ppm, and 6.94 ± 2.55 ppm, respectively. As the land cover types are mainly farmland and shrubs, the urban agglomerations of YRD, TCC, and CCG have relatively high natural emissions. Additionally, their regional CH4 sinks are affected by similar hydrothermal conditions. The high concentrations of CCG are also affected by the basin retention effect, though regional annual emissions are generally lower than those of YRD and TCC. Temporally, the increase in anthropogenic emissions in BTH from 2010 to 2021 was approximately twice that of YRD. Meanwhile, the annual XCH4 values in BTH increased at a rate of 8.95 ppb, which is slightly higher than other regions (8.61 ppb for YRD, 8.46 ppb for TCC, and 8.27 ppb for CCG). These results indicated that the regional XCH4 values calculated by the Mapping-XCH4 data accurately quantified the differences among different urban agglomerations.

4. Discussion

In the spatiotemporal kriging interpolation algorithm, the kriging variances are calculated as the mapping error. Figure 14 shows the spatiotemporal distribution of mapping uncertainties expressed by the standard deviation. In terms of space, the uncertainties in other regions are less than 14 ppb on the whole, except for regions in the Indo-China Peninsula. In the article by Sheng et al. [46], it was clearly pointed out that the uncertainty of the mapping data largely depends on the number of available data points used for spatiotemporal mapping. This can also be well demonstrated by comparing Figure 2 and Figure 14a. For the same reason, the mapping uncertainty of the coverage period of TROPOMI data is smaller than that of GOSAT. However, we found that the seasonal variation amplitude of the TROPOMI period was higher than that of the GOSAT period in the analysis of different spatial scales. In the temporal trends for different latitude zones, the difference in the 20°–30° latitude zone is more obvious than in other latitude zones.
Compared with the TCCON data, the mean difference of all the stations is greater than 5.50 ppb, and the standard deviation is greater than 10 ppb (Table 2). The Xianghe station is located in the northern part of the Beijing–Tianjin–Hebei region of China. The concentration of CH4 in the northwestern part of the station is significantly lower than that in the surrounding areas. Therefore, the mean deviation (ME) is negative, and the standard deviation is higher than at the other sites. Due to the spatial resolution of the coarse grid, the Mapping-XCH4 data are not able to accurately monitor atmospheric CH4 concentration changes at local point sources. However, for the time series, the mapping data can reflect the overall temporal variation trend of the area around the site, which can be seen in Figure 15.

5. Conclusions

Reducing CH4 emissions is recognized as an effective option for rapid climate change mitigation, especially on decadal timescales, because of its shorter lifetime than CO2. In order to fill the data gaps of individual satellite products, in this study, we mapped the temporal and spatial continuous XCH4 data (Mapping-XCH4) in monsoon regions in Asia from 2009 to 2021 using GOSAT and TROPOMI data observations. The dataset was used to analyze the long-term temporal and spatial distribution characteristics and changes.
Compared with the individual satellite observations of GOSAT and TROPOMI, the Mapping-XCH4 data reveal the regional spatiotemporal variation characteristics more comprehensively. For the spatial distribution of the multi-year average XCH4 from 2010 to 2021, there are elevated CH4 concentrations in the North China Plain, southern China, South Asia, and Southeast Asia due to the primary CH4 sources of wetlands and rice paddy fields. The low concentrations of CH4 are mainly distributed in China’s Qinghai–Tibet Plateau and the northern regions at latitudes above 40°N. The long-term trend of XCH4 from 2009 to 2021 increased from 1785.63 ppb to 1897.28 ppb, with an annual growth rate of about 8.4 ppb. In terms of seasonal changes, the CH4 concentrations in autumn and summer are generally higher than those in spring and winter because of the stronger emissions from rice paddies and wetlands in the warmer and wetter seasons. Through a spatiotemporal correlation analysis, we found that the Mapping-XCH4 data are positively correlated with emissions from the EDGAR inventory, showing a consistent spatial distribution.
These results demonstrate the potential of the mapping XCH4 dataset for monitoring long-term CH4 variations and provide new cases for further study on the driving factors of CH4 changes and the responses of CH4 concentrations observed by satellites to anthropogenic emissions at regional scales. In the future, combined with a large number of remote sensing parameters, such as surface temperature, land cover, and meteorological data, more refined interpolation will be carried out by using the deep learning method to increase the ability of the mapping data to identify point source emissions.

Author Contributions

Conceptualization, H.S., M.S. and L.L.; Data curation, H.S. and M.S.; Formal analysis, H.S., M.S. and L.L.; Methodology, H.S. and L.L.; Software, S.Z., Z.J. and K.G. 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. 2020YFA0607503, and 2022YFC3800700).

Data Availability Statement

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

Acknowledgments

We are grateful for GOSAT observations (GOSAT-XCH4) provided by Environmental Studies (NIES) from the website at (https://data2.gosat.niesgo.jp/GosatDataArchiveServ.ice/usr/download, assessed on 25 July 2022), S5P observations (S5P-XCH4) provided by Copernicus S5p Open AccessHub from the website at (https://s5phub.copernicus.eu/dhus/#/home, assessed on 25 July 2022), and the Emissions Database for Global Atmospheric Research (EGDAR)data provided by the European commission’s Joint Research Centre (JRC) at (https://edgar.jrc.ec.europa.eu/dataset_ghg70_nuts2, assessed on 29 November 2022). We thank the Total Carbon Column Observing Network (TCCON) for providing XCH4 observed data products at (https://tccondata.org/, accessed on 15 July 2022).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lin, X.; Zhang, W.; Crippa, M.; Peng, S.; Han, P.; Zeng, N.; Yu, L.; Wang, G. A Comparative Study of Anthropogenic CH4 Emissions over China Based on the Ensembles of Bottom-up Inventories. Earth Syst. Sci. Data 2021, 13, 1073–1088. [Google Scholar] [CrossRef]
  2. Shukla, P.R.; Skea, J.; Slade, R.; Al Khourdajie, A.; Van Diemen, R.; McCollum, D.; Pathak, M.; Some, S.; Vyas, P.; Fradera, R. IPCC, 2022: Climate Change 2022: Mitigation of Climate Change. Contribution of Working Group III to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University Press: Cambridge, UK; New York, NY, USA, 2022; Volume 10, p. 9781009157926. [Google Scholar]
  3. Nisbet, E.G.; Manning, M.R.; Dlugokencky, E.J.; Fisher, R.E.; Lowry, D.; Michel, S.E.; Myhre, C.L.; Platt, S.M.; Allen, G.; Bousquet, P.; et al. Very Strong Atmospheric Methane Growth in the 4 Years 2014–2017: Implications for the Paris Agreement. Glob. Biogeochem. Cycles 2019, 33, 318–342. [Google Scholar] [CrossRef] [Green Version]
  4. Wuebbles, D.J.; Hayhoe, K.A.; Kotamarthi, R. Methane in the Global Environment. Atmos. Methane Its Role Glob. Environ. 2000, 12, 304–341. [Google Scholar]
  5. World Meteorological Organization (WMO). Greenhouse Gas Bulletin (GHG Bulletin): The State of Greenhouse Gases in the Atmosphere Based on Global Observations through 2021 (No. 18|26 October 2022); WMO: Geneva, Switzerland, 2022. [Google Scholar]
  6. Bruhwiler, L.; Dlugokencky, E.; Masarie, K.; Ishizawa, M.; Andrews, A.; Miller, J.; Sweeney, C.; Tans, P.; Worthy, D. CarbonTracker-CH 4: An Assimilation System for Estimating Emissions of Atmospheric Methane. Atmos. Chem. Phys. 2014, 14, 8269–8293. [Google Scholar] [CrossRef] [Green Version]
  7. Menon, S.; Denman, K.L.; Brasseur, G.; Chidthaisong, A.; Ciais, P.; Cox, P.M.; Dickinson, R.E.; Hauglustaine, D.; Heinze, C.; Holland, E.; et al. Couplings between Changes in the Climate System and Biogeochemistry; Lawrence Berkeley National Lab: Berkeley, CA, USA, 2007. [Google Scholar]
  8. Zhang, B.; Yang, T.R.; Chen, B.; Sun, X.D. China’s Regional CH4 Emissions: Characteristics, Interregional Transfer and Mitigation Policies. Appl. Energy 2016, 184, 1184–1195. [Google Scholar] [CrossRef]
  9. Dlugokencky, E.J.; Nisbet, E.G.; Fisher, R.; Lowry, D. Global Atmospheric Methane: Budget, Changes and Dangers. Philos. Trans. R. Soc. Math. Phys. Eng. Sci. 2011, 369, 2058–2072. [Google Scholar] [CrossRef] [Green Version]
  10. Shi, T.; Han, G.; Ma, X.; Mao, H.; Chen, C.; Han, Z.; Pei, Z.; Zhang, H.; Li, S.; Gong, W. Quantifying Factory-Scale CO2/CH4 Emission Based on Mobile Measurements and EMISSION-PARTITION Model: Cases in China. Environ. Res. Lett. 2023, 18, 034028. [Google Scholar] [CrossRef]
  11. World Meteorological Organization (WMO). Greenhouse Gas Bulletin (GHG Bulletin)-No. 14: The State of Greenhouse Gases in the Atmosphere Based on Global Observations through 2017; WMO: Geneva, Switzerland, 2017. [Google Scholar]
  12. Jackson, R.B.; Saunois, M.; Bousquet, P.; Canadell, J.G.; Poulter, B.; Stavert, A.R.; Bergamaschi, P.; Niwa, Y.; Segers, A.; Tsuruta, A. Increasing Anthropogenic Methane Emissions Arise Equally from Agricultural and Fossil Fuel Sources. Environ. Res. Lett. 2020, 15, 071002. [Google Scholar] [CrossRef]
  13. 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]
  14. 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. 2013, 52, 3594–3603. [Google Scholar] [CrossRef]
  15. Xu, J.; Li, W.; Xie, H.; Wang, Y.; Wang, L.; Hu, F. Long-Term Trends and Spatiotemporal Variations in Atmospheric XCH4 over China Utilizing Satellite Observations. Atmosphere 2022, 13, 525. [Google Scholar] [CrossRef]
  16. Tomosada, M.; Kanefuji, K.; Matsumoto, Y.; Tsubaki, H. A Prediction Method of the Global Distribution Map of CO2 Column Abundance Retrieved from GOSAT Observation Derived from Ordinary Kriging. In Proceedings of the 2009 ICCAS-SICE, Fukuoka, Japan, 18–21 August 2009; IEEE: Scottsdale, AZ, USA, 2009; pp. 4869–4873. [Google Scholar]
  17. Liu, Y.; Wang, X.; Guo, M.; Tani, H. Mapping the FTS SWIR L2 Product of XCO2 and XCH4 Data from the GOSAT by the Kriging Method–a Case Study in East Asia. Int. J. Remote Sens. 2012, 33, 3004–3025. [Google Scholar] [CrossRef] [Green Version]
  18. Liu, M.; Lei, L.; Liu, D.; Zeng, Z.-C. Geostatistical Analysis of CH4 Columns over Monsoon Asia Using Five Years of GOSAT Observations. Remote Sens. 2016, 8, 361. [Google Scholar] [CrossRef] [Green Version]
  19. Li, L.; Lei, L.; Song, H.; Zeng, Z.; He, Z. Spatiotemporal Geostatistical Analysis and Global Mapping of CH4 Columns from GOSAT Observations. Remote Sens. 2022, 14, 654. [Google Scholar] [CrossRef]
  20. NIES. GOSAT Project Release Note of Bias-Corrected FTS SWIR Level 2 CO2, CH4 Products (V02.95/V02.96) for General Users; Copernicus Publications: Gottingen, Germany, 2021. [Google Scholar]
  21. Landgraf, J.; Lorente, A.; Borsdorff, T.; Hasekamp, O.P.; de Gouw, J.; Veefkind, P. Two Year of TROPOMI Methane Observations: Data Quality and Science Opportunities. In Proceedings of the AGU Fall Meeting Abstracts, San Francisco, CA, USA, 9–13 December 2019; Volume 2019, p. A53F-01. [Google Scholar]
  22. Lorente, A.; Borsdorff, T.; Butz, A.; Langerock, B.; Sha, M.K.; Hasekamp, O.P.; Landgraf, J. TROPOMI Methane Total Column Measurements from TROPOMI S5-P and Suomi-NPP: Improved Data Quality, Validation and First Applications. In Proceedings of the AGU Fall Meeting Abstracts, San Francisco, CA, USA, 9–13 December 2019; Volume 2019, p. A43J-2974. [Google Scholar]
  23. Sha, M.K.; Langerock, B.; Blavier, J.-F.L.; Blumenstock, T.; Borsdorff, T.; Buschmann, M.; Dehn, A.; De Mazière, M.; Deutscher, N.M.; Feist, D.G.; et al. Validation of Methane and Carbon Monoxide from Sentinel-5 Precursor Using TCCON and NDACC-IRWG Stations. Atmos. Meas. Technol. 2021, 14, 6249–6304. [Google Scholar] [CrossRef]
  24. Shiomi, K.; Kawakami, S.; Ohyama, H.; Arai, K.; Okumura, H.; Ikegami, H.; Usami, M. GGG2020.R0, TCCON Data from Saga (JP); Caltech Library: Pasadena, CA, USA, 2022. [Google Scholar] [CrossRef]
  25. Morino, I.; Ohyama, H.; Hori, A.; Ikegami, H. GGG2020.R0, TCCON Data from Rikubetsu (JP); Caltech Library: Pasadena, CA, USA, 2022. [Google Scholar] [CrossRef]
  26. Morino, I.; Ohyama, H.; Hori, A.; Ikegami, H. GGG2020.R0, TCCON Data from Tsukuba (JP), 125HR; Caltech Library: Pasadena, CA, USA, 2022. [Google Scholar] [CrossRef]
  27. Zhou, M.; Wang, P.; Kumps, N.; Hermans, C.; Nan, W. GGG2020.R0, TCCON Data from Xianghe, China; Caltech Library: Pasadena, CA, USA, 2022. [Google Scholar] [CrossRef]
  28. Liu, C.; Wang, W.; Sun, Y.; Shan, C. GGG2020.R0, TCCON Data from Hefei (PRC); Caltech Library: Pasadena, CA, USA, 2022. [Google Scholar] [CrossRef]
  29. Crippa, M.; Solazzo, E.; Huang, G.; Guizzardi, D.; Koffi, E.; Muntean, M.; Schieberle, C.; Friedrich, R.; Janssens-Maenhout, G. High Resolution Temporal Profiles in the Emissions Database for Global Atmospheric Research. Sci. Data 2020, 7, 121. [Google Scholar] [CrossRef] [Green Version]
  30. Muntean, M.; Janssens-Maenhout, G.; Song, S.; Giang, A.; Selin, N.E.; Zhong, H.; Zhao, Y.; Olivier, J.G.J.; Guizzardi, D.; Crippa, M.; et al. Evaluating EDGARv4.Tox2 Speciated Mercury Emissions Ex-Post Scenarios and Their Impacts on Modelled Global and Regional Wet Deposition Patterns. Atmos. Environ. 2018, 184, 56–68. [Google Scholar] [CrossRef]
  31. Han, P.; Zeng, N.; Oda, T.; Lin, X.; Crippa, M.; Guan, D.; Janssens-Maenhout, G.; Ma, X.; Liu, Z.; Shan, Y.; et al. Evaluating China’s Fossil-Fuel CO2 Emissions from a Comprehensive Dataset of Nine Inventories. Atmos. Chem. Phys. 2020, 20, 11371–11385. [Google Scholar] [CrossRef]
  32. Crippa, M.; Guizzardi, D.; Solazzo, E.; Muntean, M.; Schaaf, E.; Monforti-Ferrario, F.; Banja, M.; Olivier, J.; Grassi, G.; Rossi, S. GHG Emissions of All World Countries; Publications Office of the European Union: Luxembourg, 2021. [Google Scholar]
  33. Zeng, Z.-C.; Lei, L.; Strong, K.; Jones, D.B.; Guo, L.; Liu, M.; Deng, F.; Deutscher, N.M.; Dubey, M.K.; Griffith, D.W. Global Land Mapping of Satellite-Observed CO2 Total Columns Using Spatio-Temporal Geostatistics. Int. J. Digit. Earth 2017, 10, 426–456. [Google Scholar] [CrossRef] [Green Version]
  34. He, Z.; Lei, L.; Zhang, Y.; Sheng, M.; Wu, C.; Li, L.; Zeng, Z.-C.; Welp, L.R. Spatio-Temporal Mapping of Multi-Satellite Observed Column Atmospheric CO2 Using Precision-Weighted Kriging Method. Remote Sens. 2020, 12, 576. [Google Scholar] [CrossRef] [Green Version]
  35. Anselin, L. Local Indicators of Spatial Association—LISA. Geogr. Anal. 1995, 27, 93–115. [Google Scholar] [CrossRef]
  36. Bivand, R.S.; Wong, D.W. Comparing Implementations of Global and Local Indicators of Spatial Association. Test 2018, 27, 716–748. [Google Scholar] [CrossRef]
  37. Kennedy, J.; Eberhart, R. Particle Swarm Optimization. In Proceedings of the IEEE ICNN’95-International Conference on Neural Networks, Perth, WA, USA, 27 November–1 December 1995; Volume 4, pp. 1942–1948. [Google Scholar]
  38. Mucherino, A.; Papajorgji, P.J.; Pardalos, P.M.; Mucherino, A.; Papajorgji, P.J.; Pardalos, P.M. K-Nearest Neighbor Classification. Data Min. Agric. 2009, 34, 83–106. [Google Scholar]
  39. Huang, L.; Ning, Z.; Lin, H.; Zhou, W.; Wang, L.; Zou, J.; Xu, H. High-Pressure Sorption of Methane, Ethane, and Their Mixtures on Shales from Sichuan Basin, China. Energy Fuels 2021, 35, 3989–3999. [Google Scholar] [CrossRef]
  40. Song, K.; Zhang, G.; Yu, H.; Xu, H.; Lv, S.; Ma, J. Methane and Nitrous Oxide Emissions from a Ratoon Paddy Field in Sichuan Province, China. Eur. J. Soil Sci. 2021, 72, 1478–1491. [Google Scholar] [CrossRef]
  41. Behera, M.D.; Mudi, S.; Shome, P.; Das, P.K.; Kumar, S.; Joshi, A.; Rathore, A.; Deep, A.; Kumar, A.; Sanwariya, C. COVID-19 Slowdown Induced Improvement in Air Quality in India: Rapid Assessment Using Sentinel-5P TROPOMI Data. Geocarto Int. 2021, 37, 8127–8147. [Google Scholar] [CrossRef]
  42. Gupta, K.; Kumar, R.; Baruah, K.K.; Hazarika, S.; Karmakar, S.; Bordoloi, N. Greenhouse Gas Emission from Rice Fields: A Review from Indian Context. Environ. Sci. Pollut. Res. 2021, 28, 30551–30572. [Google Scholar] [CrossRef]
  43. Metya, A.; Datye, A.; Chakraborty, S.; Tiwari, Y.K.; Sarma, D.; Bora, A.; Gogoi, N. Diurnal and Seasonal Variability of CO2 and CH4 Concentration in a Semi-Urban Environment of Western India. Sci. Rep. 2021, 11, 2931. [Google Scholar] [CrossRef]
  44. Wu, D.; Lin, J.C.; Fasoli, B.; Oda, T.; Ye, X.; Lauvaux, T.; Yang, E.G.; Kort, E.A. A Lagrangian Approach towards Extracting Signals of Urban CO2 Emissions from Satellite Observations of Atmospheric Column CO2 (XCO2): X-Stochastic Time-Inverted Lagrangian Transport Model (“X-STILT V1”). Geosci. Model Dev. 2018, 11, 4843–4871. [Google Scholar] [CrossRef] [Green Version]
  45. Belikov, D.A.; Saitoh, N.; Patra, P.K.; Chandra, N. GOSAT CH4 Vertical Profiles over the Indian Subcontinent: Effect of a Priori and Averaging Kernels for Climate Applications. Remote Sens. 2021, 13, 1677. [Google Scholar] [CrossRef]
  46. Sheng, M.; Lei, L.; Zeng, Z.-C.; Rao, W.; Song, H.; Wu, C. Global Land 1° Mapping Dataset of XCO 2 from Satellite Observations of GOSAT and OCO-2 from 2009 to 2020. Big Earth Data 2022, 7, 170–190. [Google Scholar] [CrossRef]
Figure 1. The flowchart of generating Mapping-XCH4 data in Monsoon Asia from April 2009 to December 2021 using satellite XCH4 observations of GOSAT and TROPOMI.
Figure 1. The flowchart of generating Mapping-XCH4 data in Monsoon Asia from April 2009 to December 2021 using satellite XCH4 observations of GOSAT and TROPOMI.
Remotesensing 15 03389 g001
Figure 2. Spatial distribution of XCH4 data points after converting the spatial and temporal scales from 2019 to 2021 for (a) GOSAT and (b) TROPOMI. The legend represents the number of data points in 1° grid, which is N × 102 for GOSAT and N × 103 for TROPOMI.
Figure 2. Spatial distribution of XCH4 data points after converting the spatial and temporal scales from 2019 to 2021 for (a) GOSAT and (b) TROPOMI. The legend represents the number of data points in 1° grid, which is N × 102 for GOSAT and N × 103 for TROPOMI.
Remotesensing 15 03389 g002
Figure 3. Comparison of GOSAT-XCH4 and TROPOMI-XCH4. Scatter plot of GOSAT-XCH4 and TROPOMI-XCH4 from 2019 to 2021. X C H 4   o r i and X C H 4   u n i are described in Equation (1). The XCH4 data are averaged within 3 days and 0.1° grid.
Figure 3. Comparison of GOSAT-XCH4 and TROPOMI-XCH4. Scatter plot of GOSAT-XCH4 and TROPOMI-XCH4 from 2019 to 2021. X C H 4   o r i and X C H 4   u n i are described in Equation (1). The XCH4 data are averaged within 3 days and 0.1° grid.
Remotesensing 15 03389 g003
Figure 4. The results of XCH4 spatiotemporal correlation analysis in monsoon Asia. (a) The spatiotemporal deterministic trend of XCH4. (b) Spatiotemporal variogram models calculated by XCH4 residual component.
Figure 4. The results of XCH4 spatiotemporal correlation analysis in monsoon Asia. (a) The spatiotemporal deterministic trend of XCH4. (b) Spatiotemporal variogram models calculated by XCH4 residual component.
Remotesensing 15 03389 g004
Figure 5. Averaged mapping XCH4 in 1° latitudinal band and 3-day interval from 2009 to 2021.
Figure 5. Averaged mapping XCH4 in 1° latitudinal band and 3-day interval from 2009 to 2021.
Remotesensing 15 03389 g005
Figure 6. Time series of monthly XCH4 for different zones, (a) and its growth rates (b) from 2009 to 2021.
Figure 6. Time series of monthly XCH4 for different zones, (a) and its growth rates (b) from 2009 to 2021.
Remotesensing 15 03389 g006
Figure 7. Spatial distribution of XCH4 in the monsoon Asia from 2010 to 2021.
Figure 7. Spatial distribution of XCH4 in the monsoon Asia from 2010 to 2021.
Remotesensing 15 03389 g007
Figure 8. Seasonally averaged XCH4 calculated from Mapping-XCH4, including (a) spring (March to May); (b) summer (June to August); (c) autumn (September to November); and (d) winter (December, January, and February the next year).
Figure 8. Seasonally averaged XCH4 calculated from Mapping-XCH4, including (a) spring (March to May); (b) summer (June to August); (c) autumn (September to November); and (d) winter (December, January, and February the next year).
Remotesensing 15 03389 g008
Figure 9. Spatial associations of XCH4 in the monsoon Asia from 2010 to 2021.
Figure 9. Spatial associations of XCH4 in the monsoon Asia from 2010 to 2021.
Remotesensing 15 03389 g009
Figure 10. The clustering results of gridded XCH4 time series based on Mapping-XCH4 from 2009 to 2021. The red line represents the Heihe–Tengchong Line. The colors and the numbers represent the different clustering regions.
Figure 10. The clustering results of gridded XCH4 time series based on Mapping-XCH4 from 2009 to 2021. The red line represents the Heihe–Tengchong Line. The colors and the numbers represent the different clustering regions.
Remotesensing 15 03389 g010
Figure 11. (a) The multi-year average emissions from EDGAR during the period from 2009 to 2021, (b) the comparison of XCH4 and emissions based on clusters.
Figure 11. (a) The multi-year average emissions from EDGAR during the period from 2009 to 2021, (b) the comparison of XCH4 and emissions based on clusters.
Remotesensing 15 03389 g011
Figure 12. Spatial distribution of the four urban agglomerations in China. The background is CH4 emissions from EDGAR in Figure 11a. BTH, YRD, CCG, and TCC represent the Beijing–Tianjin–Hebei region and nearby areas, the Yangtze River Delta, the triangle of Central China, and the Chengdu–Chongqing City Group, respectively. The colors represent the multi-year average emissions from EDGAR during the period from 2009 to 2021.
Figure 12. Spatial distribution of the four urban agglomerations in China. The background is CH4 emissions from EDGAR in Figure 11a. BTH, YRD, CCG, and TCC represent the Beijing–Tianjin–Hebei region and nearby areas, the Yangtze River Delta, the triangle of Central China, and the Chengdu–Chongqing City Group, respectively. The colors represent the multi-year average emissions from EDGAR during the period from 2009 to 2021.
Remotesensing 15 03389 g012
Figure 13. The time series of XCH4 in the urban agglomerations of BTH, YRD, TCC, and CCG.
Figure 13. The time series of XCH4 in the urban agglomerations of BTH, YRD, TCC, and CCG.
Remotesensing 15 03389 g013
Figure 14. Spatial and temporal distribution of XCH4 in the monsoon Asia from 2010 to 2021 (b). The colors in (a) represents the spatiotemporal distribution of mapping uncertainties.
Figure 14. Spatial and temporal distribution of XCH4 in the monsoon Asia from 2010 to 2021 (b). The colors in (a) represents the spatiotemporal distribution of mapping uncertainties.
Remotesensing 15 03389 g014
Figure 15. Time series for matches of Mapping-XCH4 and TCCON data at Saga, Rikubetsu, Tsukuba, Xianghe, and Hefei.
Figure 15. Time series for matches of Mapping-XCH4 and TCCON data at Saga, Rikubetsu, Tsukuba, Xianghe, and Hefei.
Remotesensing 15 03389 g015
Table 1. The products of CH4 data from satellite observations.
Table 1. The products of CH4 data from satellite observations.
DatasetGOSAT-XCH4TROPOMI-XCH4
PeriodApril 2009–December 2021.November 2018–December 2021
Data ProductBias-corrected FTS SWIR L2 CH4S5P_OFFL_L2__CH4
Spatiotemporal Resolution3 days, 10.5 km1 day, ~7 km
Criteria of Data ScreeningxCH4_quality_flag = 0,
land_fraction > 90
qa_value > 0.75
Data SourceGDASCopernicus S5p Open Access Hub
Table 2. Information of Mapping-XCH4 vs. TCCON data at different sites used in this paper.
Table 2. Information of Mapping-XCH4 vs. TCCON data at different sites used in this paper.
SiteLatitudeLongitudeStart TimeEnd TimeNMEMAESTDR
Saga33.24°N130.29°E28 July 201130 June 20217478.038611.385613.35020.8970
Rikubetsu43.46°N143.77°E24 June 201430 June 202143910.824913.203112.75220.8516
Tsukuba36.05°N140.12°E28 March 201428 June 20214845.53919.018910.83860.8452
Xianghe39.8°N116.96°E14 June 201830 November 2021324–9.313313.888315.30100.7261
Hefei31.9°N119.17°E8 January 201631 December 20201858.494813.328013.78270.8256
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

Song, H.; Sheng, M.; Lei, L.; Guo, K.; Zhang, S.; Ji, Z. Spatial and Temporal Variations of Atmospheric CH4 in Monsoon Asia Detected by Satellite Observations of GOSAT and TROPOMI. Remote Sens. 2023, 15, 3389. https://doi.org/10.3390/rs15133389

AMA Style

Song H, Sheng M, Lei L, Guo K, Zhang S, Ji Z. Spatial and Temporal Variations of Atmospheric CH4 in Monsoon Asia Detected by Satellite Observations of GOSAT and TROPOMI. Remote Sensing. 2023; 15(13):3389. https://doi.org/10.3390/rs15133389

Chicago/Turabian Style

Song, Hao, Mengya Sheng, Liping Lei, Kaiyuan Guo, Shaoqing Zhang, and Zhanghui Ji. 2023. "Spatial and Temporal Variations of Atmospheric CH4 in Monsoon Asia Detected by Satellite Observations of GOSAT and TROPOMI" Remote Sensing 15, no. 13: 3389. https://doi.org/10.3390/rs15133389

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