Next Article in Journal
Adverse Health Effects (Bronchitis Cases) Due to Particulate Matter Exposure: A Twenty-Year Scenario Analysis for the Greater Athens Area (Greece) Using the AirQ+ Model
Previous Article in Journal
Evaluation of Eutrophication in Jiaozhou Bay via Water Color Parameters Determination with UAV-Borne Hyperspectral Imagery
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Spatial-Resolution Methane Emissions Calculation Using TROPOMI Data by a Divergence Method

1
School of Environment and Spatial Informatics, China University of Mining and Technology, Xuzhou 221116, China
2
Geophysical Survey Team of Hebei Province Coal Geological Bureau, Xingtai 054000, China
3
Jiangsu Key Laboratory of Coal-Based Greenhouse Gas Control and Utilization, China University of Mining and Technology, Xuzhou 221008, China
*
Author to whom correspondence should be addressed.
Atmosphere 2023, 14(2), 388; https://doi.org/10.3390/atmos14020388
Submission received: 20 January 2023 / Revised: 9 February 2023 / Accepted: 13 February 2023 / Published: 16 February 2023
(This article belongs to the Section Air Quality)

Abstract

:
Methane (CH4) is the second-largest greenhouse gas emitted by human activity and natural sources after carbon dioxide (CO2). Its relatively short lifetime in the atmosphere (about 12 years) means that we can mitigate the human impacts of climate change in a relatively short period of time by reducing CH4 emissions. The creation of CH4 emissions management policies can be based on the distribution maps of surface CH4 concentration that are in large-scale and at high-resolution. The estimate of CH4 emissions with broad coverage are provided by currently extensively used satellite data supplemented with data from model simulations. However, it is at low spatial resolution. In this paper, through the combination of atmospheric CH4 observations from the TROPOMI sensor and wind data from the ECMWF global reanalysis, a straightforward divergence method is proposed to estimate the surface CH4 emissions in China from March 2019 to September 2022 at a resolution of 7 km × 7 km. This method was compared with the average annual CH4 emissions of Emissions Database for Global Atmospheric Research (EDGARv7.0), and the Root Mean Square Error (RMSE) is 2.53 kg/km2/h and within error envelop (EE) is 72.93%, which represents the proportion of reliable values under certain uncertain conditions. We estimated that the average annual CH4 emissions in China from 2019 to 2022 is 81 Tg, with the lowest emissions in 2021 (75 Tg) due to the impact of COVID-19. In 2021, the largest anthropogenic emissions in China are from agriculture, energy activities and livestock, accounting for 28% (20.8 Tg), 25% (18.9 Tg) and 19% (13.9 Tg) of total emissions, respectively, while wetlands, as the largest natural source, produce 14% (10.5 Tg) of CH4 emissions.

1. Introduction

Methane (CH4), which makes up more than 20% of all greenhouse gases (GHGs) from both natural and anthropogenic sources, is the second most prevalent GHGs after carbon dioxide (CO2) [1,2]. The World Meteorological Organization’s most recent study of observation data from the Global Atmospheric Watch Program reveals that atmospheric CH4 concentration has climbed from 722 ppb (part per billion) in 1750 to 1908 ± 2 ppb in 2021, an increase of 18 ppb with respect to the previous year and the record is constantly increasing [3]. In addition, the radiative forcing from atmospheric CH4 is 0.48 ± 0.05 W m−2 and makes up 17% of the global radiative forcing of GHGs [4]. Due to significant global warming potential and brief atmospheric lifetime (approximately 12 years) of CH4, reducing its emissions can partially alleviate the effects of climate change on humans in a short amount of time. China is one of the largest CH4 emissions in the world, and will actively reduce CH4 emissions in the next decade [5]. Plans for reducing CH4 emissions will be more effective if they take into account the precise temporal and spatial variations of surface emissions.
Major sources of CH4 emissions in China include agriculture, energy activities, livestock, waste and wetlands. To effectively and accurately monitor CH4 emissions, we need detailed analysis and monitoring of major sources. Agriculture includes agricultural waste burning and agricultural soils. Agricultural waste burning is mainly the burning of straw of rice, corn and other crops, while the water content of agricultural waste is generally high, burning agricultural waste will also produce a large amount of CH4. CH4 emissions from agriculture are mainly caused by rice paddy soil because of warm, moist paddy soils providing ideal conditions for CH4 production [6]. Energy activities include industry, transport, energy for buildings, fuel exploitation [7], iron and steel production, etc. CH4 is a large part of the world’s energy, used to heat homes, cook food, heat water, generate electricity and even power transportation. Industries that contribute to anthropogenic CH4 emissions are mainly the Oil refineries and transformation industry, power plants, manufacturing plants and steel mills, whose combustion converts CH4 energy into carbon dioxide for release into the atmosphere. Cattle contribute the majority of greenhouse gas emissions from Enteric fermentation, followed by other ruminants such as buffalo, sheep and goats. Enteric fermentation occurs when anaerobic microorganisms, called methanogens, break down and ferment food present in the animal’s digestive tract to produce compounds that are then absorbed by the animal host. This digestive process allows ruminants to eat more plant material that would otherwise be indigestible. About 7–10% of ruminants’ energy is lost in the form of CH4 and excreted by enteric fermentation through burping [8]. In 2020, the waste sector, municipal solid waste and wastewater, accounted for nearly 20 percent of all human-related CH4 emissions, the third largest global source of CH4 emissions after enteric fermentation and energy activities [9]. Wetlands are one of the largest natural sources of CH4-producing emissions globally, which are estimated to make up 20% to 30% of the global CH4 emissions budget [10]. In wetlands, CH4 is produced by microbial activity and can be transported from the soil to the atmosphere through a variety of methods. According to the government data of the National Forestry and Grassland Administration, China’s wetland area is about 56.35 million hectares, accounting for 4% of the world’s wetland area. Thirteen cities in China have been elected to the Convention on Wetlands, making it the country with the largest number of international wetland cities in the world. The NASA Carbon Monitoring System (CMS) program describes, quantifies, understands and predicts the evolution of global carbon sources and sinks through satellite observations and modeling analysis capabilities, providing estimates of monthly CH4 emissions from global wetlands.
The two primary techniques used to measure atmospheric CH4 concentration currently are bottom-up ground-based methods [11,12] and top-down satellite platforms [13,14,15,16,17]. The advantage of satellite-based monitoring is that it can identify the characteristics of temporal and spatial CH4 concentration variations on a local, regional, and global scale. The Atmospheric Infrared Sounder (AIRS) on EOS/Aqua satellite has been used to analyze the distribution of CH4 in the middle and upper troposphere of China (MUT-CH4) from 2003 to 2008 [14]. The results show that the growth in China is generally more significant than in other parts of the world. Based on CH4 vertical columns measured by the Scanning Imaging Absorption spectroMeter for Atmospheric CHartographY (SCIAMACHY) onboard Envisat, Zhao et al. presented temporal and spatial variations of the three-order trapezoid spatial gradient of CH4 in China from 2003 to 2005 (0.5° latitude × 0.5° longitude) [18]. Miller et al. used CH4 observations from the Greenhouse gases Observing SATellite (GOSAT) for aggregate 2.0° × 2.5° latitude-longitude grid boxes between September 2009 and September 2015 to show the latest trends in Asia’s total anthropogenic and natural emissions, with a particular focus on China, largely attributed to coal mining [19]. Qin et al. obtained the spatial-temporal characteristics of seasonal variation of atmospheric CH4 concentration over China from 2003 to 2021 using the CH4 column Homogeneous Dry Air Mole Fraction (XCH4) product of SCIAMACHY, GOSAT and the Sentinel 5 Precursor (S5P) [20]. Since its October 2017 launch, the TROPOspheric Monitoring Instrument (TROPOMI) on board the S5P satellite has measured CH4 concentrations at an unprecedented 7 km × 7 km resolution (upgraded to 5.5 km × 7 km in August 2019), outperforming previous widely used instruments such as GOSAT and SCIAMACHY. Chen et al. quantify CH4 emissions in China and the contributions from different sectors by inverse analysis of 2019 TROPOMI satellite observations of atmospheric CH4 at 0.25° × 0.3125° resolution [21].
However, previous studies have shown that (1) the spatial resolution of previous satellite data is too low, resulting in low coverage of CH4 emissions data; (2) The model’s simulation data and satellite data are combined to cover the entire study area, but the spatial resolution of the results is lower than that of the satellite itself. In this study, a straightforward divergence method is used successfully to estimate surface CH4 emissions from TROPOMI satellite data, and the spatial resolution is consistent with the satellite data.

2. Data and Methodology

2.1. Data

2.1.1. Surface Atmospheric XCH4 from TROPOMI

The S5P is the first of the atmospheric composition Sentinels, launched on 13 October 2017. For this study, we used S5P OFFO(L2) CH4 products from March 2019 to September 2022 with a spatial resolution of 7 km × 7 km, as shown in Table 1. Focusing on the XCH4 bias corrected product, the accuracy offered in the data product (defined as the standard deviation of the retrieval noise, which describes the impact of the measurement noise on the retrieval) and the quality descriptor (“qa_value”) is suitable for data use. The “qa_value” indicates the status and quality of the retrieval output. To assure that the highest quality data is used, pixels that are classified with qa_value < 0.5 were filtered to ensure data quality. The entire atmospheric column was divided into 12 layers in the TROPOMI XCH4 retrieval, and the lowest layer is used to represent the surface atmospheric column concentration. We used the mole content of CH4 in atmosphere bottom layer in INPUT_DATA divided by the mole content of dry air in atmosphere bottom layer to obtain the surface column averaged dry air mixing ratio of CH4 (surface atmospheric XCH4).

2.1.2. Wind Data

ERA5 is the fifth generation ECMWF reanalysis for global climate and weather, which uses the laws of physics to combine model data with observations from across the world to form a globally complete and consistent dataset [22]. ERA5 has been re-gridded to a regular latitude-longitude grid of 0.25 degrees for the reanalysis. Eastward wind, 10 U parameter from ECMWF (grib variable 165), is the horizontal component of the wind at 10-meter height in the eastward direction. Northward wind, 10 V parameter from ECMWF (grib variable 166), is the horizontal component of the wind at 10-meter height in the northward direction, whose data format is NC FLOAT. The TROPOMI orbit timestamp, which is generally one hour before the Riyadh overpass, was interpolated to using a linear interpolation from the six-hourly output.

2.1.3. CH4 Emissions from EDGAR

The gridded CH4 emissions database in China was from the Emissions Database for Global Atmospheric Research (EDGAR), which is a widely used inventory for anthropogenic emissions of greenhouse gases on Earth. EDGAR provides CH4 emission estimation from sector-specific grid maps at 0.1° × 0.1° resolution with yearly, monthly, and hourly data using the IPCC methodology [23,24]. This study collected data on yearly anthropogenic CH4 emissions from all sectors for 2019–2021, and we divided them into Agricultural, energy activities, livestock and Waste.

2.1.4. Wetland CH4 Emissions

The CMS program aims to make a significant contribution to characterizing, quantifying, understanding, and predicting the evolution of global carbon sources and sinks by improving the monitoring of carbon storage and fluxes. WetCHARTs v1.3.1 provides global monthly wetland CH4 emissions estimates at 0.5 by 0.5-degree resolution for the period 2001–2019, whose model drivers are replaced from using ERA-interim to ERA5 reanalysis data [25]. This product is the intended use is as a process of information wetlands CH4 emissions data set, used for atmospheric chemistry and transport modeling.

2.2. Methodology

2.2.1. Surface Atmospheric XCH4 Model

The TROPOMI CH4 data product is given in the form of total column-averaged dry-air mole fraction, XCH4. It is calculated from the CH4 vertical sub-column elements x i and the dry-air column V a i r , d r y calculated with meteorology input from ECMWF:
XCH 4 = i = 0 n x i V a i r , d r y
The entire atmospheric column was divided into 12 layers in the TROPOMI CH4 profile data and the bottom layer represents the concentration of the surface atmospheric column, whose average height is about 700 m. In order to obtain surface XCH4 quickly, we obtain atmospheric CH4 a priori vertical profile data and dry air density column from TROPOMI official supporting data. To assure that the highest quality data is used, pixels that are classified with qa_value < 0.5 were filtered to ensure data quality. XCH4 in surface is derived as follow:
XCH 4 = V C H 4 V a i r
V C H 4 and V a i r stand for the bottom layer of the TROPOMI CH4 profile data and dry air density used by the retrieval of TROPOMI XCH4, respectively.

2.2.2. A Divergence Method to Quantify CH4 Emissions

The divergence can be used to describe the degree of divergence of vector field at different points in space. According to the continuity equation for steady state, the divergence (D), emission (E) and sink (S) can be expressed as: D = E + S [26]. Because CH4 is a long-lived gas that can exist in the atmosphere for 12.5 years, there is atmospheric CH4 emission and background at the surface. The sink term can be ignored and the background (B) concentration is assumed to be completely homogeneous [27]. As D is a linear operator, the relationship between D, E and B can be expressed as: E = D − B and the average emissions over time can be calculated.
Here, we extract top-down CH4 emission maps on high spatial resolution based on TROPOMI CH4 columns V combined with wind fields ω from the European Center for Medium Range Weather Forecasts (ECMWF). The CH4 fluxes is given as F = V ω and the divergence D works on horizontal fluxes (F): D = F , thus E = F = ( V ω ) , where ω is divided into vector eastward wind ( u ) and northward wind ( v ) at 10 m height.
In this study, numerical derivatives for D were calculated as the second-order central difference. For each grid cell, its neighboring grid cells along south-north ( v component) and east-west ( u component) directions are firstly used to obtain the divergence. However, it is hard to know the exact background, so we use the regional background to approximate the background of surface XCH4, that is defined as the average of the lower 10 percentile of its surrounding ±5 grid cells (11 × 11 = 121 grid cells in total by taking the current grid cell as the center). So, the daily variations of CH4 emission can be written as:
E = ( ( V     B ) ω )
where V stands for column of CH4 in the surface and B stands for divergence of the regional background.

2.2.3. Validation by EDGAR Anthropogenic CH4 Emissions

To evaluate the feasibility of our approach, anthropogenic CH4 emission data from EDGARv7.0 were used to validate estimated CH4 emission data we calculated through the divergence method. EDGAR reports on the evolution of national CH4 emission inventories around the world over time and provides a 0.1° × 0.1° grid representing CH4 emission sources. EDGARv7.0 provides CH4 emissions per sector and country from 1970 to 2021 and grid maps are expressed in kg substance /km2/h for the network Common Data Form (NetCDF) files.
Using the modified divergence method, daily CH4 emissions in 2021 were estimated, and their annual mean values were compared with EDGAR’s 2021 annual emissions, totaling 10,709 measurements. Due to the complex terrain, we choose the eastern plain as the verification target. Figure 1 shows the scatter plot between our estimated CH4 emissions and EDGAR’s anthropogenic CH4 emissions, with the root mean square error (RMSE) of 2.53 kg/km2/h and within error envelop (EE) of 72.93%, indicating that our annual estimation is slightly higher than EDGAR’s due to uncounted anthropogenic and natural CH4 emissions in EDGAR’s emissions and the error of satellite-based background concentration.

3. Results

3.1. Spatial Distribution Patterns of Surface Atmospheric XCH4 across China

In order to analyze the surface spatial distribution patterns of XCH4 in China, the surface spatial distribution of multi-year-averaged XCH4 concentrations from S5p from January 2019 to September 2022 is shown in Figure 2. The calculation details are provided in Section 2.2.1. The range and average of XCH4 concentrations from S5P were 1900 to 2150, and 1998 ppb. The XCH4 concentrations showed greatly spatial variation in different regions of China and presented spatial cluster in different regions. Overall, there were low XCH4 concentrations in southwest China, high concentrations in east China and moderate concentrations in the north. Because CH4 is a long-lived gas that can exist in the atmosphere for 12.5 years, the spatial variations of atmospheric CH4 concentrations might reflect the accumulation of CH4 emissions and transportations.
According to the available data (from Jan. 2019 to Sept. 2022, qa_value > 0.5), the high concentration areas in China mainly include central China, East China, the Beijing-Tianjin-Hebei region, the Sichuan Basin and Part of the northern part of Xinjiang Uygur Autonomous Region, with an average concentration of 2064 ppb, which is 66 ppb higher than the China’s average; The areas with consistently low concentrations are Tibet and Qinghai Province, with an average concentration of 1934 ppb, which are 64 ppb lower than the China’s average; The rest of China was in the middle, with an average of 1975 ppb, 23 ppb lower than the China’s average.

3.2. Analysis of Divergence Method of Calculating CH4 Emissions in China

In order to analyze the spatial and temporal distribution of CH4 emissions in China, the divergence method is used to calculate China’s CH4 emissions through TROPOMI surface XCH4 concentrations from March 2019 to September 2022, as shown in Figure 3. Excluding the missing data, the regions with high emission rates are the Tibetan Plateau, parts of Yunnan, parts of northern Xinjiang, the Sichuan Basin, central China, eastern China, and the three provinces in northeast China. China’s northwest and west region has virtually no CH4 emissions, according to EDGAR CH4 emissions data. Although these areas do not have significant anthropogenic CH4 emissions, they have a relatively high surface CH4 concentration levels and relatively high wind speeds and change of wind direction throughout the years (Figure 3). In the Tibetan Plateau, the daily amount of TROPOMI secondary data is sufficient, but it is very sparse with a large number of vacancy values. The divergence method is used to estimate CH4 emissions on a daily basis, which leads to the abnormal divergence (vacancy value is set to zero, resulting in the value points identified as high divergence points). Similarly, in Yunnan Province, not only is the daily data very sparse, but also the amount of data is very small, resulting in high divergence. In the hilly areas of southern China, there are also large data missing, resulting in the loss of computing continuity. In addition, in areas with large altitude changes, such as Kunlun Mountains, Qilian mountains, Hengduan Mountains, Greater Khingan Mountains and hilly areas in southern China, the data quality is poor, resulting in abnormal divergence calculation (positive or negative values are both too high). Wind speed and wind direction are also the most important factors affecting divergence. Large divergence errors will be caused at the interfaces of different terranes (where there are high wind speeds in opposite directions), such as China’s First Ladder and the second ladder, as well as China’s second ladder and the third ladder (Kunlun Mountains and Greater Khingan Mountains, which typically connect the Qinghai-Tibet Plateau and the Tarim Basin).

3.3. CH4 Emissions in China

Due to missing data, for each grid cell, the divergence is calculated only when there are valid values along the north-south and east-west directions. Excessive wind speed will affect the conservation of mass in the region, so the wind speed is set to be less than 10 m/s. A negative value represents a reduction in CH4 concentration in the area, indicating that more CH4 was transported than emitted over a period of years.
Based on daily TROPOMI observation from March 2019 to September 2022, we estimated multi-year averages CH4 emissions in the Chinese region with a resolution of 7 km × 7 km. The calculation details are provided in Section 2.2.2. When the divergence of CH4 is obtained, CH4 emissions can be obtained by averaging them over time. As shown in Figure 4a,b, TROPOMI observations are mainly distributed in the northern, central and eastern regions of China, where CH4 emissions can be better estimated. In addition, in central, eastern and northern China, where a small amount of complex terrain exists, daily data volumes are plentiful and of higher quality than in other regions, resulting in smoother and more robust performance on averages. Since the daily data of the Qinghai-Tibet Plateau is very sparse and there is almost no anthropogenic CH4 emissions, we choose to mask this part. The accumulation of divergence over time manifests as local CH4 emissions. We estimated that the average annual CH4 emissions in China is 81 Tg y−1, which is higher to EDGARv7.0 average annual CH4 anthropogenic emissions from 2019 to 2021 (66 Tg y−1), which can be due to the uncertainty of satellite observation and the existence of natural sources. The average annual CH4 emissions in China from 2019 to 2022 are 78 Tg, 87 Tg, 75 Tg, 85 Tg respectively, and the emissions in 2021 due to the epidemic were less than the previous year.
To analyze the differences in CH4 emissions between northern and southern China, we compared the CH4 emissions characteristics of the Beijing-Tianjin-Hebei region and Jiangsu Province in 2021. The terrain in the Beijing-Tianjin-Hebei region is very different in height, being high in the northwest and low in the southeast. Figure 4c shows that the distribution characteristics of CH4 emissions in the Beijing-Tianjin-Hebei region are completely different from those in Jiangsu Province. The ratio of plateau, plain, and mountainous land in Hebei Province is 1:4:5, which blocks the air flow in the west and north and makes it difficult for haze pollution to spread. Due to the influence of mountainous terrain, a large number of data gaps exist in the northern part of Hebei Province. Hebei province is rich in mineral resources, with large numbers of coal mines and coal-fired power plants in the cities of Shijiazhuang, Xingtai, and Handan, causing a lot of pollution. Jiangsu Province is abundant in water resources and has the highest percentage of plain land in China, up to 86.9%. Since paddy fields are the primary crop, agriculture accounts for the majority of Jiangsu Province’s CH4 emissions. Two significant freshwater lakes in Jiangsu Province, Taihu Lake and Hongze Lake, have substantial CH4 emissions levels close by, which may be brought on by the interaction of rice fields and natural wetlands. In 2021, We estimated that Hebei and Jiangsu Province emitted 6.2 Tg and 3.8 Tg of CH4 annually, respectively. CH4 emissions from energy activities in Hebei Province are higher than those from paddy fields and natural wetlands in Jiangsu Province.
EDGARv7.0 average annual CH4 anthropogenic emission was divided into 22 sectors in 2021, which we summarize into four sectors, including agriculture, energy activities, livestock andwaste, as shown in Figure 5a–d.
The Sichuan Basin, southern and northeastern China are large areas where rice is grown. Soils in the south generally contain more moisture than those in the north, and the entire Yangtze River basin, in particular, is the worst emitter of CH4. The most serious areas of agriculture waste burning are southern Guangdong Province, Guangxi Zhuang Autonomous Region and Taiwan Province. Our estimate for agriculture in 2021 is about 20.8 Tg, the highest anthropogenic CH4 emissions of any industry. Anthropogenic CH4 emissions from energy for buildings, like those from most factories, are distributed in and around provincial capitals, largely because of increasing urbanization. CH4 emissions from road transport such as rail and road transport occur on major national and provincial highways across the country. Important fuel exploitation activities caused more than 80 percent of CH4 emissions in Shanxi Province, with a smaller amount in Henan and Shandong provinces. Anthropogenic CH4 emissions from energy activities have a characteristic national distribution, totaling about 18.9 Tg in 2021, but are mainly concentrated in central and eastern China. Nationwide, CH4 emissions from enteric fermentation are mainly distributed in Jiangsu, Anhui, Shandong and Henan regions, followed by southwest China, and a small amount of CH4 emissions exists in northeast China, with a total estimated CH4 emission of 13.9 Tg in 2021. In China, CH4 emissions generated by waste in all cities are estimated to be 11.7 Tg y−1. Due to China’s large population and increasing purchasing power, a large amount of consumption-related waste is accumulated in major cities across the country, especially in super-large cities such as Beijing, Shanghai, Guangzhou and Chengdu. According to the data, the estimated amount of China’s wetland CH4 emissions is 10.5 Tg y−1, mainly distributed in central China, including Hubei, Anhui and Jiangxi provinces, as shown in Figure 5e.

4. Conclusions

CH4 is the second most important greenhouse gas causing greenhouse effect, with a short lifespan of about 12 years. Therefore, effective monitoring of CH4 emissions can provide a basis for the government’s carbon emission reduction policies. In order to obtain high resolution surface CH4 emission data in China, a simple divergence method is proposed in this paper. Firstly, the surface CH4 column concentration can be obtained by combining the CH4 profile data of TROPOMI with the dry air density column. The precise background concentration was then replaced with a regional CH4 concentration of less than 10 percent of XCH4 in a regional grid; Finally, using the surface CH4 column concentration of TROPOMI and wind data of ECMWF, the divergence method is used to estimate surface CH4 emissions.
CH4 emissions calculated by divergence method were evaluated against CH4 emissions data from EDGAR, showing a good result in the overall estimation. However, there is a phenomenon of high CH4 concentration but few emissions in complex topographic areas, so the estimation of this method is a certain overestimated.
The surface CH4 column concentrations distribution shows that the high concentration of CH4 is in eastern China and the low concentration is in southwest China. Based on the divergence method, the average annual surface estimated CH4 emissions in China from 2019 to 2022 were 78 Tg, 87 Tg, 75 Tg and 85 Tg, respectively. Affected by the pandemic, the average annual estimated CH4 emissions in 2021 were significantly reduced. In 2021, agriculture, energy activities and livestock are the three sectors with the largest anthropogenic estimated CH4 emissions in China, which emit 20.8 Tg, 18.9 Tg and 13.9 Tg respectively, while wetland, the largest natural source of CH4 emissions, releases 10.5 Tg, accounting for 14% of the total estimated emissions.
The advantage of the method proposed in this paper is that the high resolution and wide coverage data of TROPOMI can be used to estimate surface CH4 emissions in a large range. However, CH4 emissions are overestimated in areas with sparse TROPOMI data and areas with complex terrain, so we will improve this method in the future.

Author Contributions

Conceptualization, Y.X., C.W. and S.L.; methodology, Y.X., C.W.; software, S.L.; validation, S.L., P.G., L.Z., C.J. and B.Z.; formal analysis, S.L. and B.H.; investigation, S.L., L.Z. and P.G.; resources, S.L.; data curation, S.L. and L.Z.; writing—original draft preparation, S.L.; writing—review and editing, S.L.; visualization, S.L.; supervision, Y.X. and C.W.; project administration, Y.X. and C.W.; funding acquisition, Y.X. and C.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by the National Natural Science Foundation of China (NSFC) under Grant No. 42275147 and the Project Support of Hebei Province under Grant 13000022P00B04410020E.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank all the principal investigators and their staff for providing the data and products used in this investigation. The TROPOMI data and ERA5 hourly data were obtained from the ESA Copernicus Open Access Hub and the European Centre for Medium-Range Weather Forecasts (ECWMF). The EDGARv7.0 CH4 emissions data were obtained from Emissions Database for Global Atmospheric Research (EDGAR). The WetCHARTs v1.3.1 data were obtained from the NASA Carbon Monitoring System (CMS) program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shindell, D.; Kuylenstierna, J.C.; Vignati, E.; van Dingenen, R.; Amann, M.; Klimont, Z.; Anenberg, S.C.; Muller, N.; Janssens-Maenhout, G.; Raes, F.; et al. Simultaneously mitigating near-term climate change and improving human health and food security. Science 2012, 335, 183–189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Saunois, M.; Bousquet, P.; Poulter, B.; Peregon, A.; Ciais, P.; Canadell, J.G.; Dlugokencky, E.J.; Etiope, G.; Bastviken, D.; Houweling, S.; et al. The global methane budget 2000–2012. Earth Syst. Sci. Data 2016, 8, 697–751. [Google Scholar] [CrossRef] [Green Version]
  3. World Meteorological Organization. WMO Greenhouse Gas Bulletin (GHG Bulletin)—No.18: The State of Greenhouse Gases in the Atmosphere Based on Global Observations through 2021; World Meteorological Organization: Geneva, Switzerland, 2022. [Google Scholar]
  4. Alexander, L.; Allen, S.; Bindoff, N.; Breon, F.-M.; Church, J.; Cubasch, U.; Emori, S.; Forster, P.; Friedlingstein, P.; Gillett, N.; et al. Climate Change 2013: The Physical Science Basis, in Contribution of Working Group I (WGI) to the Fifth Assessment Report (AR5) of the Intergovernmental Panel on Climate Change (IPCC); IPCC: Geneva, Switzerland, 2013. [Google Scholar]
  5. Zhang, Y.; Fang, S.; Chen, J.; Lin, Y.; Chen, Y.; Liang, R.; Jiang, K.; Parker, R.J.; Boesch, H.; Steinbacher, M.; et al. Observed changes in China’s methane emissions linked to policy drivers. Proc. Natl. Acad. Sci. USA 2022, 119, e2202742119. [Google Scholar] [CrossRef] [PubMed]
  6. Kumar, J.I.N.; Viyol, S.V. Short-term diurnal and temporal measurement of methane emission in relation to organic carbon, phosphate and sulphate content of two rice fields of central Gujarat, India. Paddy Water Environ. 2009, 7, 11–16. [Google Scholar] [CrossRef]
  7. Singh, A.K.; Kumar, J. Fugitive methane emissions from Indian coal mining and handling activities: Estimates, mitigation and opportunities for its utilization to generate clean energy. Energy Procedia 2016, 90, 336–348. [Google Scholar] [CrossRef]
  8. Gao, Z.L.; Yuan, H.J.; Ma, W.Q.; Liu, X.J.; Desjardins, R.L. Methane emissions from a dairy feedlot during the fall and winter seasons in Northern China. Environ. Pollut. 2011, 159, 1183–1189. [Google Scholar] [CrossRef] [PubMed]
  9. Ghosh, P.; Shah, G.; Chandra, R.; Sahota, S.; Kumar, H.; Vijay, V.K.; Thakur, I.S. Assessment of methane emissions and energy recovery potential from the municipal solid waste landfills of Delhi, India. Bioresour. Technol. 2019, 272, 611–615. [Google Scholar] [CrossRef] [PubMed]
  10. Koffi, E.N.; Bergamaschi, P.; Alkama, R.; Cescatti, A. An observation-constrained assessment of the climate sensitivity and future trajectories of wetland methane emissions. Sci. Adv. 2020, 6, eaay4444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Liu, L.X.; Zhou, L.X.; Wen, M.; Zhang, F.; Fang, S.X.; Yao, B. Characteristics of Atmospheric CH4 Concentration Variations at Four National Baseline Observatories in China. Adv. Clim. Chang. Res. 2009, 5, 285–290. [Google Scholar]
  12. Fang, S.-X.; Zhou, L.-X.; Xu, L.; Yao, B.; Liu, L.-X.; Xia, L.-J.; Wang, H.-Y. CH4 concentrations and the variation characteristics at the four WMO/GAW background stations in China. Huan Jing Ke Xue Huanjing Kexue 2012, 33, 2917–2923. [Google Scholar] [PubMed]
  13. Xiong, X.; Houweling, S.; Wei, J.; Maddy, E.; Sun, F.; Barnet, C. Methane plume over south Asia during the monsoon season: Satellite observation and model simulation. Atmos. Chem. Phys. 2009, 9, 783–794. [Google Scholar] [CrossRef] [Green Version]
  14. Zhang, X.Y.; Bai, W.G.; Zhang, P.; Wang, W.H. Spatiotemporal variations in mid-upper tropospheric methane over China from satellite observations. Chin. Sci. Bull. 2011, 56, 3321–3327. [Google Scholar] [CrossRef] [Green Version]
  15. Zhang, X.Y.; Jiang, H.; Wang, Y.Q.; Han, Y.; Buchwitz, M.; Schneising, O.; Burrows, J.P. Spatial variations of atmospheric methane concentrations in China. Int. J. Remote Sens. 2011, 32, 833–847. [Google Scholar] [CrossRef]
  16. Wu, X.D.; Zhang, X.Y.; Chuai, X.W.; Huang, X.J.; Wang, Z. Long-Term Trends of Atmospheric CH4 Concentration across China from 2002 to 2016. Remote Sens. 2019, 11, 538. [Google Scholar] [CrossRef] [Green Version]
  17. Zhang, G.L.; Xiao, X.M.; Dong, J.W.; Xin, F.F.; Zhang, Y.; Qin, Y.W.; Doughty, R.B.; Moore, B. Fingerprint of rice paddies in spatial-temporal dynamics of atmospheric methane concentration in monsoon Asia. Nat. Commun. 2020, 11, 554. [Google Scholar] [CrossRef] [Green Version]
  18. Zhao, H.Q.; Zhang, L.F.; Wu, T.X.; Duan, Y.N.; Cen, Y. Analysis on the Spatial-Temporal Variations of Methane over China Using Sciamachy Data. In Proceedings of the 2013 IEEE International Geoscience and Remote Sensing Symposium—IGARSS, Melbourne, VIC, Australia, 21–26 July 2013; pp. 1786–1789. [Google Scholar] [CrossRef]
  19. Miller, S.M.; Michalak, A.M.; Detmers, R.G.; Hasekamp, O.P.; Bruhwiler, L.M.P.; Schwietzke, S. China’s coal mine methane regulations have not curbed growing emissions. Nat. Commun. 2019, 10, 303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Qin, J.Y.; Zhang, X.Y.; Zhang, L.J.; Cheng, M.M.; Lu, X.H. Spatiotemporal Variations of XCH4 across China during 2003-2021 Based on Observations from Multiple Satellites. Atmosphere 2022, 13, 1362. [Google Scholar] [CrossRef]
  21. Chen, Z.C.; Jacob, D.J.; Nesser, H.; Sulprizio, M.P.; Lorente, A.; Varon, D.J.; Lu, X.; Shen, L.; Qu, Z.; Penn, E.; et al. Methane emissions from China: A high-resolution inversion of TROPOMI satellite observations. Atmos. Chem. Phys. 2022, 22, 10809–10826. [Google Scholar] [CrossRef]
  22. The European State of the Climate 2017 Was Compiled by the Copernicus Services of the European Centre for Medium-Range Weather Forecasts (ECMWF)—The Climate Change Service (C3S) and the Atmosphere Monitoring Service (CAMS). Available online: https://cds.climate.copernicus.eu/cdsapp#!/home (accessed on 19 January 2023).
  23. Crippa, M.; Guizzardi, D.; Solazzo, E.; Muntean, M.; Schaaf, E.; Monforti-Ferrario, F.; Banja, M.; Olivier, J.; Grassi, G.; Rossi, S.; et al. GHG Emissions of All World: 2021 Report; Publications Office of the European Union: Brussels, Belgium, 2021. [Google Scholar]
  24. Crippa, M.; Guizzardi, D.; Banja, M.; Solazzo, E.; Muntean, M.; Schaaf, E.; Pagani, F.; Monforti, F.; Olivier, J.; Quadrelli, R.; et al. CO2 Emissions of All World Countries. JRC/IEA/PBL 2022 Report; Publications Office of the European Union: Luxembourg, 2022. [Google Scholar]
  25. Bloom, A.A.; Bowman, K.W.; Lee, M.; Turner, A.J.; Schroeder, R.; Worden, J.R.; Weidner, R.J.; McDonald, K.C.; Jacob, D.J. CMS: Global 0.5-Deg Wetland Methane Emissions and Uncertainty (WetCHARTs v1.3.1); NASA: Washington, DC, USA, 2021. [Google Scholar] [CrossRef]
  26. Beirle, S.; Borger, C.; Dorner, S.; Li, A.; Hu, Z.K.; Liu, F.; Wang, Y.; Wagner, T. Pinpointing nitrogen oxide emissions from space. Sci. Adv. 2019, 5, eaax9800. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Liu, M.Y.; van der, A.R.; van Weele, M.; Eskes, H.; Lu, X.; Veefkind, P.; de Laat, J.; Kong, H.; Wang, J.X.; Sun, J.Y.T.; et al. A New Divergence Method to Quantify Methane Emissions Using Observations of Sentinel-5P TROPOMI. Geophys. Res. Lett. 2021, 48, e2021GL094151. [Google Scholar] [CrossRef]
Figure 1. Scatter plots between EDGAR emissions higher than 0.0 kg/km2/h and estimated CH4 emissions. The red line defines an expected error, which characterizes the estimated CH4 emissions as being three times larger than the EDGAR emissions under certain uncertain conditions.
Figure 1. Scatter plots between EDGAR emissions higher than 0.0 kg/km2/h and estimated CH4 emissions. The red line defines an expected error, which characterizes the estimated CH4 emissions as being three times larger than the EDGAR emissions under certain uncertain conditions.
Atmosphere 14 00388 g001
Figure 2. Surface spatial distribution of multi-year averages of atmospheric XCH4 over China from January 2019 to September 2022.
Figure 2. Surface spatial distribution of multi-year averages of atmospheric XCH4 over China from January 2019 to September 2022.
Atmosphere 14 00388 g002
Figure 3. (a) Surface spatial distribution of multi-year averages CH4 of divergence over China from January 2019 to September 2022. (b) The elevation map that is generated from GMTED2010 data set. (c) Multi-year averages distribution of eastward wind in China from 2019 to 2022. (d) Multi-year averages distribution of northward wind in China from 2019 to 2022.
Figure 3. (a) Surface spatial distribution of multi-year averages CH4 of divergence over China from January 2019 to September 2022. (b) The elevation map that is generated from GMTED2010 data set. (c) Multi-year averages distribution of eastward wind in China from 2019 to 2022. (d) Multi-year averages distribution of northward wind in China from 2019 to 2022.
Atmosphere 14 00388 g003
Figure 4. (a) Surface spatial distribution of averages CH4 of divergence over China from March 2019 to September 2022. (b) Number of observations on that grid over China from March 2019 to September 2022. (c) The estimated CH4 emissions are based on (a).
Figure 4. (a) Surface spatial distribution of averages CH4 of divergence over China from March 2019 to September 2022. (b) Number of observations on that grid over China from March 2019 to September 2022. (c) The estimated CH4 emissions are based on (a).
Atmosphere 14 00388 g004aAtmosphere 14 00388 g004b
Figure 5. Spatial distribution of CH4 emissions from five source sectors in China.
Figure 5. Spatial distribution of CH4 emissions from five source sectors in China.
Atmosphere 14 00388 g005
Table 1. Description of data used in the paper.
Table 1. Description of data used in the paper.
DatasetParameter(s)ResolutionTime RangeDownload Link
TROPOMI CH4Priori CH4 profile7 km × 7 km2019.03–2022.09https://s5phub.copernicus.eu/dhus/#/home accessed on 8 October 2022.
Dry air columns
ECMWF Wind10 m u-component of wind0.25° × 0.25°2019.03–2022.09https://cds.climate.copernicus.eu/ accessed on 1 October 2022.
10 m v-component of wind
EDGAR CH4CH4 emissions0.1° × 0.1°2019–2021https://edgar.jrc.ec.europa.eu/ accessed on 1 August 2022.
WetCHARTs v1.3.1Wetlands CH4 emissions0.5° × 0.5°2019https://doi.org/10.3334/ORNLDAAC/1915 accessed on 1 October 2022.
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

Li, S.; Wang, C.; Gao, P.; Zhao, B.; Jin, C.; Zhao, L.; He, B.; Xue, Y. High-Spatial-Resolution Methane Emissions Calculation Using TROPOMI Data by a Divergence Method. Atmosphere 2023, 14, 388. https://doi.org/10.3390/atmos14020388

AMA Style

Li S, Wang C, Gao P, Zhao B, Jin C, Zhao L, He B, Xue Y. High-Spatial-Resolution Methane Emissions Calculation Using TROPOMI Data by a Divergence Method. Atmosphere. 2023; 14(2):388. https://doi.org/10.3390/atmos14020388

Chicago/Turabian Style

Li, Shengwei, Chunbo Wang, Pengyuan Gao, Bingjie Zhao, Chunlin Jin, Liang Zhao, Botao He, and Yong Xue. 2023. "High-Spatial-Resolution Methane Emissions Calculation Using TROPOMI Data by a Divergence Method" Atmosphere 14, no. 2: 388. https://doi.org/10.3390/atmos14020388

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