Next Article in Journal
Australian Bushfires (2019–2020): Aerosol Optical Properties and Radiative Forcing
Next Article in Special Issue
Impact of Collaborative Agglomeration of Manufacturing and Producer Services on Air Quality: Evidence from the Emission Reduction of PM2.5, NOx and SO2 in China
Previous Article in Journal
Skill and Intercomparison of Global Climate Models in Simulating Wind Speed, and Future Changes in Wind Speed over South Asian Domain
Previous Article in Special Issue
Inventory of Commercial Cooking Activities and Emissions in a Typical Urban Area in Greece
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Modification of Fraser’s Method for the Atmospheric CO2 Mass Estimation by Using Satellite Data

Department of Industrial Engineering, University of Bologna, 40136 Bologna, Italy
*
Author to whom correspondence should be addressed.
Atmosphere 2022, 13(6), 866; https://doi.org/10.3390/atmos13060866
Submission received: 18 March 2022 / Revised: 18 May 2022 / Accepted: 21 May 2022 / Published: 25 May 2022

Abstract

:
One of the most critical greenhouse gases in the atmosphere is carbon dioxide (CO2) due to its long-lasting and negative impact on climate change. The global atmospheric monthly mean CO2 concentration is currently greater than 410 ppm which has changed dramatically since the industrial era. To choose suitable climate change mitigation and adaptation strategies it is necessary to define carbon dioxide mass distribution and global atmospheric carbon dioxide mass. The available method to estimate the global atmospheric CO2 mass was proposed in 1980. In this study, to increase the accuracy of the available method, various observation platforms such as ground-based stations, ground-based tall towers, aircrafts, balloons, ships, and satellites are compared to define the best available observations, considering the temporal and spatial resolution. In the method proposed in this study, satellite observations (OCO2 data), from January 2019 to December 2021, are used to estimate atmospheric CO2 mass. The global atmospheric CO2 mass is estimated around 3.24 × 1015 kg in 2021. For the sake of comparison, global atmospheric CO2 mass was estimated by Fraser’s method using NOAA data for the mentioned study period. The proposed methodology in this study estimated slightly greater amounts of CO2 in comparison to Fraser’s method. This comparison resulted in 1.23% and 0.15% maximum and average difference, respectively, between the proposed method and Fraser’s method. The proposed method can be used to estimate the required capacity of systems for carbon capturing and can be applied to smaller districts to find the most critical locations in the world to plan for climate change mitigation and adaptation.

1. Introduction

Most of the literature agrees that greenhouse gases (GHGs) trap heat in the atmosphere and lead to global warming. One of the most critical GHGs is carbon dioxide (CO2) which has attracted more attention than other GHGs due to its long-lasting presence in the atmosphere and negative impact on climate change. Furthermore, always increasing data would suggest that CO2 concentration increases year after year. If 20 parts per million (ppm) was assumed based on high accuracy Antarctic ice-core records in 6000 Before Common Era (BCE) [1], more than ten times greater concentrations, i.e., up to around 281 ppm, were recorded during the Industrial Revolution between the 17th and the 18th centuries [2,3]. The global monthly mean CO2 concentration is currently greater than 410 ppm, according to the National Oceanic and Atmospheric Administration (NOAA) website [4]. However, no simple correlation exists between global warming and available CO2 concentration data. For this purpose, new CO2 indexes should be introduced for the investigation of global warming problems. In fact, CO2 concentration results from a complex balance between sources and sinks, such as, for example, anthropogenic activities and natural phenomena [5,6,7]. Therefore, many parameters must be considered, as shown in Table A2 in the Appendix A section.
However, despite the efforts to measure CO2 concentration, the CO2 mass calculation approach is inadequate. For instance, a preliminary estimation of about 7.15 × 1011 tons was reported in 1980 [8], while a more recent publication of the Global Carbon Budget in 2019 estimates an amount up to 8.60 × 1011 tons [9]. Therefore, 1.45 × 1011 tons of CO2 seem to have been emitted in approximately 39 years, resulting in a mean annual positive flux to the atmosphere of 3.7 Gton/year (=1.45 × 1011/39). The annual growth rate in the atmospheric CO2 mass was calculated from the concentration data reported by Dlugokencky and Tans [4], i.e., from the Global Greenhouse Gas Reference Network (GGGRN) that currently consists of 84 active sites in 37 countries. The complete GGGRN sites’ list is reported in Table A3 in the Appendix A section. Despite the number of GGGRN sites increasing through the years with a maximum number of active sites up to 116 in 2011 (see Figure A1 in the Appendix A), the currently active 85 observatory sites appear insufficient to estimate the global CO2 mass.
The first improvement in CO2 mass estimation would be using data from more datasets and not limited to the one currently used. In fact, according to Jiang and Yung [10], many CO2 datasets exist even if not included in the algorithms used for mass estimation. The list of existing surface-based and aircraft-based CO2 concentration datasets is reported below:
Ground-based CO2 concentration observations:
  • The Earth System Research Laboratory of the National Oceanic and Atmospheric Administration (NOAA ESRL) [11];
  • Total Carbon Column Observing Network (TCCON) [12].
Airborne based CO2 concentration observations:
  • The Earth System Research Laboratory of the National Oceanic and Atmospheric Administration (NOAA ESRL) [11];
  • Comprehensive Observation Network for Trace Gases by Airliner (CONTRAIL) [13];
  • Intercontinental Chemical Transport Experiment-North America (INTEX-NA) [14];
  • High-performance Instrumented Airborne Platform for Environmental Research Pole-to-Pole Observations (HIPPO) [15];
  • In-Service Aircraft for a Global Observing System (IAGOS) [16];
  • Carbon in Arctic Reservoirs Vulnerability Experiment (CARVE) [17].
In addition to ground-based sites and airborne observation, data sets from ships can be acquired using research ships or commercial ones. For example, the NOAA National Centers for Environmental Information (NCEI) hosts a data management project, the Ocean Carbon Data System (OCADS), where data from deep and shallow waters are recorded to analyze CO2 fluxes between the atmosphere and oceans. The following list includes ship-based observation under the OCADS project:
  • Ships Of Opportunity Program (SOOP) Data [18];
  • Global Surface pCO2 (LDEO) [19].
In addition, there are atmospheric CO2 concentration data sets generated in the open ocean sites using a moored autonomous system [20].
Although many algorithms exist at the present state of the art to measure CO2 fluxes and concentration, they are too complex to give an answer to several research and industrial questions, limiting their potential applications. To give an accurate estimation and location of CO2 mass amount through the use of simple algorithms is a research gap to be filled. For this purpose, i.e., to apply simple algorithms avoiding the use of state-of-the-art algorithms, big data for CO2 are required. Since it is not economically possible to build new measuring sites around the world, the elaboration of the data coming from satellites is the best alternative. Particularly, the advantages and the challenges of CO2 satellites monitoring for climate governance and for applications at national/regional, megacity and point source levels was already reviewed by [21]. For this purpose, the following satellite datasets can be considered for CO2 concentration observations:
  • The Orbiting Carbon Observatory-2 (OCO-3) [22];
  • The Orbiting Carbon Observatory-2 (OCO-2) [23];
  • The Greenhouse Gases Observing Satellite (GOSAT) [24];
  • Thermal Emission Sounder (TES), measurement instrument installed on Aura satellite by NASA [25];
  • Atmospheric InfraRed Sounder-Aqua satellite (AIRS) [26];
  • The Infrared Atmospheric Sounding Interferometer (IASI) is an instrument flown on METOP satellite [27];
  • The Atmospheric Chemistry Experiment (ACE) SciSat [28];
  • The scanning imaging absorption spectrometer for atmospheric cartography (SCIAMACHY) onboard the Environmental Satellite (ENVISAT) [29];
  • Other satellite missions are launched or will be launched in the future, such as Tansat, Carbonsat, MERLIN, Sentinel-5p, MicroCarb, and ASCENDS [3], which would increase the amount of available information.
Since more and more data will be available in the future from satellites, a new methodology for the atmospheric CO2 mass estimation is described in the paper. To the best knowledge of the authors there has not been any study that uses solely satellite data to calculate global atmospheric CO2 mass; however, there are some studies based on satellite data and simulation such as the atmospheric chemical transport model. In particular, the proposed approach ensures the estimation of the global mass but also the analysis of the atmospheric CO2 mass variation in a specific location. Therefore, the new method makes it possible to perform a comparative analysis of the atmospheric CO2 mass for different locations and time periods. The results of the proposed method can be used to estimate the required capacity of systems for carbon capturing based on the CO2 mass. In addition, since the methodology is based on the division of the Earth’s surface into smaller cells using satellite resolution it can be applied to smaller districts to find the most critical locations in the world to plan properly for climate change mitigation and adaptation.
The purpose of this paper is to validate a model using the best available observation platforms considering the resolution, coverage, and accuracy of the data to define the global atmospheric carbon dioxide mass regarding the mass distribution on the Earth with higher precision in comparison to the Fraser method. Since satellites were considered as input for the methodology, the validation of their data with respect to other platforms introduced in the Materials and Method section, i.e., ground-based stations, airborne and ships observations, is reported in the Results and Discussion sections. Finally, the new methodology for atmospheric CO2 mass described in the Materials and Method section is validated with state-of-the-art values and discussed in the Results and Discussion section.

2. Materials and Method

2.1. Simplified Fraser’s Method for the Calculation of CO2 Atmospheric Mass

According to Fraser et al. [8], the global atmospheric carbon content was 7.15 × 1014 kg in 1980. The authors suggest Figure 1 as a schematic representation of the complex procedure which was used in [11] to estimate the global atmospheric carbon dioxide mass. The validation of the proposed methodology can be performed by comparing the results using the following steps. As shown, the first step is the calculation of the number of dry air moles (block c) in the atmosphere by dividing the global atmospheric mass of dry air (block a) by dry air mean molecular weight (block b) according to Equation (1):
Air dry , mol = Air dry , mass Air dry , mw
where:
Airdry,mol = Dry air moles (mol)
Airdry,mass = Global atmospheric mass of dry air (kg), that is 5.12 × 1018 kg;
Airdry,mw = Dry air mean molecular weight (kg/mol), that is 0.02897 kg/mol.
In the second step, the number of CO2 moles in the atmosphere is calculated (block e). For this purpose, the average global CO2 concentration (block d) is multiplied by the number of dry air moles (block c) according to Equation (2):
CO 2 , mol = CO 2 , con , avg , vol × Air dry , mol × 10 6
where:
CO2,mol = CO2 moles (mol)
CO2,con,avg,mol = Average global CO2 concentration (ppmv—part per million by volume)
The average global CO2 concentration is calculated as the ratio between the CO2 and the dry air as in Equation (3):
CO 2 , con , avg , vol = CO 2 , vol Air dry , vol × 10 6
where:
CO2,vol = The volume of carbon dioxide (m3)
Airdry,vol = The volume of dry air (m3)
Since the desired parameter is the number of CO2 moles, Equation (3) is manipulated through the ideal gas law to obtain Equation (4), then Equation (2) is achieved by manipulating Equation (4):
CO 2 , con , avg , vol = CO 2 , mol Air dry , mol × 10 6
The last step is the calculation of the global atmospheric CO2 mass (block g). The value is calculated by Equation (5) as the product between the number of CO2 moles in the atmosphere (block e) and the CO2 molecular weight (block f).
CO 2 , gac = CO 2 , mol × CO 2 , mw
where:
CO2,gac = The global atmospheric carbon dioxide content (kg)
CO2,mw = CO2 molecular weight (kg), that is 44.01 × 10−3 kg/kmol.
Equation (5) is elaborated with the use of Equation (2) and Equation (1) to obtain Equation (6):
CO 2 , gac = CO 2 , con , avg , vol × Air dry , mass Air dry , mw × CO 2 , mw × 10 6

2.2. Data Sources for the Comparison of Existing Platforms for CO2 Concentration’s Measurement

Fraser’s study used only 21 sites to calibrate the proposed model for CO2 atmospheric concentration. However, through the years, more platforms have been established to measure CO2 concentration, such as ground-based platforms (including ground-based stations and tall towers), aircraft, balloons, ships and satellite-based observations.
Ground-based measurement is a fixed-space measurement of CO2 in a specific location on the earth. In addition to the earth coordinates (latitude and longitude), the elevation of the sampling vessel, i.e., the height regarding the level of the sea, is also fixed. Ground-based stations are research stations located at remote sites such as, but not limited to, islands, mountains, and coasts, although, some of them are located within a short distance from cities to assess the urban greenhouse gas emission. The selection of the sites is justified since the air samples taken at these locations would be easier to be integrate in global transport models. The Mauna Loa observatory site, located on the north flank of Mauna Loa Volcano in the main island of Hawaii (19.54° N, 155.58° W, 3397 m above the sea level (a.s.l.)) was the first station where the CO2 concentration was measured [30]. Mauna Loa is one of the so far called “Baseline Observatories” of the NOAA network, i.e., representative of the background air for a large region unaffected by local sources of pollution. In addition to Mauna Loa, observatories located at Barrow (Alaska), American Samoa, and the South Pole belong to the baseline stations. However, other ground-based stations are currently operating. Among the ground-based stations, the mentioned baseline stations and TCCON stations seem to be more reliable according to the literature [31]. The list of the worldwide ground-based stations is reported in Table A5 in the Appendix A. Tall towers are used since the 1990s to estimate the vertical CO2 concentration gradient in continental areas and to minimize the impact of local sources and sinks [32]. Thanks to tall towers the impact of remote and local emission sources can be taken into account [33]. However, since the building of a new tall tower costs many millions of dollars [34], the realization of new towers to increase existing spatial resolution is economically infeasible. The list of the worldwide ground-based tall towers in the NOAA Earth System Research Laboratory’s Global Greenhouse Gas Reference Network is reported in Table A6 in the Appendix A.
Airborne measurement has been considered in several studies as a method to collect CO2 concentration data and to validate the satellite and ground-based measurements based on their high precision. Several means of aerial transport such as aircraft [35], helicopters [36] and balloons [37] can be utilized, and the measurement can be performed by flask sampling and/or in-situ methods. As reported by [38], the measurement of CO2 concentration in air samples from aircraft began in 1957 at the Institute of Meteorology in Stockholm, where a specific program was performed until 1961 to take air samples at 1000 m a.s.l. Airborne measurement by means of aircraft is limited to the height of flight and the path, while the measurement by means of helicopter and balloons can provide samples at the various desired heights (vertical profile), and times, because of their ability to fly vertically; however, it will be affected by the capability of the helicopter and balloon. Concerning measurement, since it is undertaken during the flight, only one measurement for each set of spatial coordinates is allowed in case of aircraft observation. Therefore, more than one flight is required to take different measurements in a specific location. Furthermore, the sampling time interval depends on the type of observation, such as in-situ or flask sampling, and also the travel duration. In some cases, the sampling with flask is performed on the return path in order to minimize the time between measurements and the analysis, since it can affect the accuracy of the results.
Another way of collecting samples is the use of ships to cover relatively wide regions. The problem with this measurement is the necessity of a large number of ships to cover the waters in the whole world; in addition, very long times are required to cover all the world’s water surfaces. However, this method is relatively cheap and affordable. The sampling time interval and the sampling location depends on the program or the ship which is used, e.g., some sampling flasks or analyzers are installed onboard the commercial ships, and the path and times of the observation follow the defined path and time table of the journey, while the rest are installed on research ships, making it possible to plan sampling. Because of the limitation in the time intervals and sampling path of this platform, a detailed assessment of this platform will not be carried out in this study.
Due to low coverage of the airborne and ground observation, it is difficult to measure the CO2 concentration globally; therefore, the utilization of satellite retrieval was recommended by several authors, especially for the areas with a low density of observing stations. According to Yanfang Hou et al. [39], the Scanning Imaging Absorption Spectrometer for Atmospheric Cartography (SCIAMACHY) on board ENVISAT, which was launched in 2002, is the first satellite instrument with the aim of CO2 measurement in the lowest atmospheric layers, i.e., up to 50 km a.s.l. However, the first CO2 concentration in the upper troposphere (less than 20 km), using satellite, was retrieved by the Advanced Earth Observing Satellite (ADEOS) in 1996, by using the Interferometric Monitor for Greenhouse Gases (IMG) [40]. Satellite observations have some advantages in comparison to other means of observation, such as high coverage, but there are several challenges such as, for example, but not limited to, the accuracy of the data, data filtration (e.g., in case of the presence of clouds), and the life span of the satellite (for example, the mission duration of ADEOS satellite was less than a year). In addition, the method which is used in satellites is different from the other methods, e.g., it is not necessary to have flask samples, which increases the pace of data observation.
The most common method for the validation of data retrieved by means of satellites is comparison with calibrated ground-based data. OCO2 and ground-based comparison was made by Wunch et al. [41], Bi et al. [42], Timofeyev et al. [43], Wu et al. [44], O’Dell et al. [45], Liang et al. [46], and Liang A. [47], considering various factors such as modes of observation, satellite data version and different bands, i.e., the channel which the OCO2 satellite measures the sunlight backscattered by the Earth’s surface and atmosphere. GOSAT and ground-based observations were compared by Qin et al. [48], Dan-dan et al. [49], Velazco et al. [50], Eguchi et al. [51], Ohyama et al. [52], Rokotyan et al. [53], Yates et al. [54], Qu et al. [55], Zeng et al. [56], and Wunch et al. [57]. In many papers, several satellites were simultaneously compared with the ground-based data. Yuan et al. [58] compared the data of in-situ measurement and satellite ones, SCIAMACHY on ENVISAT, TANSO-FTS on GOSAT, and OCO2. Buchwitz et al. [59,60] during the GHG-CCI project (Climate Change Initiative (CCI)), compared TCCON data with data from SCIAMACHY/ENVISAT and TANSO/GOSAT satellites. Miao et al. [61] compared GOSAT, SCIAMACHY and AIRS with TCCON, finding that the AIRS data perform better in coverage and accuracy than the two others, in the case of the monthly mean. Avelino and Arellano [62] validated the data from AIRS, GOSAT at mid-atmosphere with ground stations. Zhang et al. [63,64] made a comparison of the data from ground stations with AIRS, SCIAMACHY, and GOSAT. Jiang et al. [65] made comparison between GOSAT, TES, AIRS, and TCCON. Reuter et al. [66] and Michael Buchwitz et al. [67] studied the difference between SCIAMACHY/ENVISAT, TANSO-FTS/GOSAT, and TCCON. Dils et al. [68] compared the TCCON data with various algorithms in different satellites and found a precision of around 2.4–2.5 ppm for almost all algorithms. Various algorithms in GOSAT were investigated by Kim et al. [69], Wunch et al. [31], Dongxu Yang et al. [70], and Lindqvist et al. [71]. These are a part of studies that have been done on difference between satellite observations and ground-based results. The available data on concentration differences of satellites and TCCON stations in these studies are illustrated in Section 3.
Several studies compared data retrieved from satellites with those from airborne measurement. However, since no measurement is possible with aircraft at altitude higher than 15 km a.s.l., stratospheric balloons are used in the range 15–35 km a.s.l. [72]. Tadic and Michalak [73] compared the data from aircraft, GOSAT and OCO2 satellites and found that the difference could be over 0.5 ppmv between aircraft and satellites. Maddy et al. [74] made a comparison among all available data from NOAA ESRL/GMD aircraft and AIRS during 2005, finding an agreement of around 0.5%. Chahine et al. (2005) [75] compared the annual cycle from AIRS and CONTRAIL over the western Pacific and found a good agreement with AIRS in both hemispheres. Uspensky et al. [76] applied an improved scheme for XCO2 on AIRS data in Siberia and cloud-cleared IASI data, comparing it with the YAK-AEROSIB aircraft campaign, finding an error of 2.2 ppmv. In another study, Kukharskii and Uspenskii [77] worked on a numerical solution for the XCO2 data retrieved from AIRS and compared it with airborne data over the areas of boreal forests (the Novosibirsk region) and ecosystems (the region of Surgut), resulting in an error no worse than 1%. Frankenberg et al. [78] compared the data from HIPPO flights with GOSAT, TES, and AIRS concluding that over the remote Pacific Ocean the GOSAT satellite, with about 0.5 ppm accuracy, has the best performance among the 3 assessed satellites.
To investigate the difference between the measurement of CO2 concentration by satellites and other platforms, i.e., the error of the data, the average value, and the standard deviation reported by each study was considered. To have an aggregate value, since the number of samples in each of the referenced studies is different, the average mean and the pooled standard deviation is used to combine all data sets.
The average mean of CO2 concentration difference between satellites and other platforms is calculated as in Equation (7):
μ p = i = 1 N n i . μ i N
where:
µp = The average mean difference between satellites and other platforms CO2 concentration
µi = Mean difference between satellites and other platforms CO2 concentration in each data set
ni = Number of measurements (samples) per data set
N = Total number of measurements
The pooled standard deviation value is calculated by Equation (8) as suggested by [79]:
σ p = 1 N K i = 1 N n i 1 . σ i 2
where:
σp = Pooled standard deviation
σi = Standard deviation in each data set
K = Number of data sets

2.3. The New Proposed Model for Atmospheric CO2 Mass Calculation

The whole Earth surface is considered as the area of study, as the main purpose of this study is to validate a method for the calculation of the atmospheric carbon dioxide mass. The main influential factors to define the global atmospheric mass of CO2 are (i) its concentration, (ii) total dry air mass, and (iii) air molecular weight. Considering the limited number of stations and observations in Fraser et al. [8] and the uncertainties in the listed factors, a new approach to calculate CO2 mass is designed. The methodology is pictured in the block diagram of Figure 2. As shown, the main difference with respect to Fraser’s methodology is that the new method uses higher resolution data for CO2 concentration and is not limited to values calculated by elaborating the data coming from a few stations located around the world (block a).
Information from satellite observation was used to allocate to each cell. For this purpose, OCO2 and OCO3 satellites were considered, since they are the latest satellites launched for the carbon dioxide observation goal. OCO2 measures atmospheric carbon dioxide [80] from the Earth surface up to the satellite [81] by means of spectrometers using the reflected sunlight intensity from CO2 in a column of air instead of direct measurement. Wavelength bands which are measured by OCO2 are 0.765 μm, related to oxygen, and two CO2 bands at 1.61 μm and 2.06 μm [82]. Diffraction grating is used to separate the reflected light energy into a spectrum of multiple component colours [83].
OCO2 was launched on 2 July 2014, and it orbits at 705 km elevation; it captures approximately 1 million soundings each day, of which around 10% are aerosols/cloud free which could be used to measure XCO2 [81]. Temporal resolution of OCO2 is 16 days and its spatial resolution is 2.25 km × 1.29 km [80].
Considering the satellite daily orbit coverage, it was decided to use 4-day data as the complete surface coverage. As shown in Figure A3, in the Appendix A, for OCO2, the satellite takes almost four days to cover the entire Earth surface. Furthermore, it is assumed that the change of the global CO2 atmospheric mass is negligible in four days. Therefore, the 4-day observation was chosen as a representative to calculate the CO2 global atmospheric mass.
Bias-corrected data from both satellites were assessed to check the applicability of the data. OCO2 and OCO3 satellites’ bias-corrected files, version 9r, were taken from the NASA website [80]. The data were in the netCDF format. R programming language was used to extract the desired data, namely, the longitude, latitude, quality flag, XCO2 (that is the column average carbon dioxide concentration in each cell), total water vapor column, and surface pressure. In particular, the quality flag defines if the data are acceptable or not. Data with a quality flag equal to 1 was not accepted in accordance with OCO2 and OCO3 documentation [84] and not considered in the following elaborations. Based on a preliminary analysis, it was decided to use only data from OCO2. In fact, OCO3 was characterized by a low number of observations that passed the quality flag filtration, around 40,000 observations in average for each day of 4-day observation. Furthermore, OCO3 data were available only since August 2019, and data were missing for several days.
A computational grid compatible with the data’s resolution was designed to divide the Earth’s surface into cells (block b). Since the Earth’s surface is 514,720,000 km2 and each cell has an area of 2.90 km2 (=2.25 km × 1.29 km), that is the spatial resolution of the satellites used in the calculation; approximately 177,500,000 cells (514,720,000/2.90) are considered.
The following steps (from block c to block i) are essentially the same as Fraser’s methodology but they are applied to each cell. Codes were written in MATLAB for the elaboration of the data while the available “curve fitting” was applied to do regression. It should be noted that to calculate the atmospheric mass of dry air in each cell (block c), data about the total surface pressure and the column of water vapor was taken from the NASA website [80]. Through Dalton’s Law, the dry air pressure in each cell was calculated. Thus, the dry air mass, which can be used in Equation (1), can be calculated, considering the cell area and gravity.
The final step is the summation of the cell values to calculate the global CO2 atmospheric mass by Equation (9):
CO 2 , gac = i = 1 N c CO 2 , cac , i 2.25 × 1.29   N c × S earth
where:
CO2,cac,I = CO2 mass in the i-th cell
Nc = number of cells
Searth = Earth surface, i.e., 510.1 × 106 km2
To validate the proposed model, Fraser’s model [8] was applied. Assuming the CO2 concentration data from NOAA, the global CO2 atmospheric mass was calculated for the period from 1980 to 2021. However, since the elaboration of the model is time-consuming, the global CO2 atmospheric mass from OCO2 was calculated for 2019 and 2020.

3. Results and Discussion

3.1. Validation of the Simplified Fraser’s Method for the Calculation of CO2 Atmospheric Mass

The authors calculated the global amount of CO2 mass by assuming an average CO2 atmospheric concentration of 337.04 ppmv, based on a two-dimensional global atmospheric CO2 transport model calibrated through the data from 21 stations [85]. To calibrate the model, the authors calculated the difference between the CO2 annual mean concentration in each station, with the South Pole station taken as reference.
Based on Equation (6), and the average global concentration calculated by Fraser et al., the amount of global CO2 mass was calculated as shown in Equation (10). More than 2.6 × 1015 kg of CO2 resulted by the application of the methodology proposed by the authors
CO 2 , gac = 337.04 × 5.12 × 10 18 0.02897 × 10 6 × 44.01 × 10 3 = 2.62 × 10 15 kg
The authors calculated also the global carbon mass equal to 7.15 × 1014 kg, simply by the substitution of the CO2 molecular weight with carbon molecular weight in Equation (6).
According to Equation (6), the average global carbon dioxide concentration, the global atmospheric mass of dry air, and dry air mean molecular weight affect CO2 mass estimation. The authors declared a probable uncertainty of 0.5 to 1% relating to carbon mass calculation divided as:
-
0.4–0.9% related to CO2 concentration. Only 21 stations were considered to design and calibrate the model as representative of the entire world. Furthermore, no stations were located in Europe, Asia, Africa, and South America, resulting in unbalanced earth surface coverage. In addition, the number of observations in which the annual mean CO2 concentration is calculated is not provided in Fraser’s study. Therefore, it is not possible to evaluate if the reported values cover all diurnal, daily, weekly, monthly, and seasonal changes or not.
-
0.1% for the air global atmospheric mass and the air mean molecular weight. Concerning the global atmospheric mass of dry air, a value equal to 5.12 × 1018 kg was used in the model. However, in 1994 a more accurate estimation of around 5.132 × 1018 kg was given by Trenberth and Guillemot [86], while, in 2005, Trenberth and Smith [87] estimated the dry air mass as 5.1352 ± 0.0003 × 1018 kg. With respect to Fraser’s model, an error equal to 0.29% results. Concerning the dry air average molecular weight, the effect of boundary conditions such as temperature and humidity is not taken into account.
-
By comparing the reported global atmospheric carbon dioxide mass in Fraser’s study and the result achieved by applying the information from this study in the Equation (6), it can be assumed that the procedure used in Fraser’s study is the same as Figure 1. Considering the complexity of Fraser’s method, the paper shows how it is possible to work on a simpler methodology with similar or even higher accuracy.

3.2. Assessment and Comparison of Existing Platforms

Existing platforms are compared through the definition of suitable Key Performance Indicators (KPI), as shown in Table 1. In this table, “ground-based stations”, “aircraft, helicopter and balloon”, “satellites” and “ships” are compared based on the available literature at the state of the art, including measurement accuracy, precision, coverage, and time. Based on the measurements described in the literature, as shown in Table 1, the accuracy of the data is almost the same, less than 0.5 ppm, except for ships where different instruments were used by researchers resulting in a wide interval of accuracy.
Almost all platforms could have high accuracy and precision, and the distinctive indices which make a difference between the platforms seem to be the coverage and the time scale required for observation. As can be seen, for the global observation the choice could be use of satellite data. However, it is possible to use other platforms and global atmospheric transport models to estimate the global atmospheric carbon dioxide concentration.
The temporal resolution of satellites provided in this table is related to the satellites which are operating, and the data available. The temporal resolution of the satellites is higher compared to the other platforms if global coverage is desired.
Since the variation in the carbon dioxide concentration related to the surface sinks and sources are typically less than 1 ppm, and seasonal and annual XCO2 variation are small in comparison to the mean atmospheric concentration, thus, 1–2 ppm precision is needed for satellite retrievals [88].

3.3. Comparison of CO2 Concentration Measured by Satellites with TCCON

Figure A2 illustrates the comparison of satellite observations and TCCON stations based on the data provided in Table A4. Since satellites’ observations are compared in Table 1 to show the accuracy and reliability, various satellites are compared with TCCON in this figure. By using the mentioned formula, and the data available in Table A4 in the Appendix A, it is possible to calculate the average mean and pooled standard deviation of the datasets for which the number of observations is provided in the context of the papers. As can be seen in this figure, the mean difference range is between −2 and 2 ppm and mean ± standard deviation ranges between −5 and 4 ppm. According to Equations (7) and (8), the mean average of the comparison between satellites and TCCON is −0.08 ppm and the pooled standard deviation is ±1.66 ppm.

3.4. Global Atmospheric CO2 Mass Calculation

The application of Fraser’s methodology to the NOAA data gives the results shown in Figure A4 in the Appendix A. As shown, a continuously increasing trend occurs in the period. The atmospheric CO2 mass increased up to 3.22 × 1015 kg in 2021, corresponding to a yearly increase of around 1.44 × 1010 tons per year between 1980 and 2021.
The proposed model was applied to OCO2 data, resulting in Figure 3 for the period 1 January 2019 to 31 November 2020. The yellow line is the global CO2 atmospheric mass calculated using the data of the satellite based on the methodology proposed in this paper. The blue line is a 12th polynomial regression curve for the global atmospheric CO2 mass calculated by the satellite (R2 = 0.71). The dark line is the curve designed from the data obtained by NOAA measurements applying Fraser’s methodology, (R2 = 0.86). Since the polyline regression with high degrees results in huge anomalies in the boundary of data the last part is neglected, and the assumed trend is provided, which is shown in the dashed red line. The available CO2 mean concentrations reported in NOAA database are weekly and monthly, of which the monthly one is used in this study. It can be seen that there is less variation in the curve related to NOAA database which goes back to the difference between the time frames.
As shown by the trend, the CO2 mass reaches its peak approximately between March and May, while the minimum occurs in the period between August and November. An upward seasonal trend of global atmospheric CO2 mass appears. The seasonal changes are probably due to various natural and anthropogenic parameters, which are different between the NH and SH. During spring and the beginning of summer in the NH, global atmospheric CO2 mass reaches its annual maximum. At the same time, the minimum happens at the end of summer and during the NH’s autumn.
Figure 3 shows the 2019–2021 period; the same cyclic trend appears in the global mass calculation using Fraser’s method and the proposed methodology. In this figure, a good agreement appears between the two methodologies. In fact, a slight difference exists between Fraser’s methodology using the NOAA dataset and the proposed methodology. By comparing the fitting curve and black curve it can be seen that the results are almost the same except for extremums, and May-July is the period with the best consistency of results.
Table 2 summarizes the differences during the period 2019–2020. As shown, a slightly greater amount of CO2 was calculated with the proposed methodology. A maximum difference of 1.25% was calculated for August 2019, while for 2020 the maximum difference was around 0.3%. Finally, an average difference of 0.15% was calculated for the entire investigating period. These differences could be due to the assumptions made in Fraser’s method, a lower number of observations and the amount of total water vapor column and surface pressure in each cell to calculate dry air moles for each cell instead of the whole Earth. Since both methods give estimations of atmospheric carbon dioxide mass it is not possible to determine which method is more accurate. However, given the use of a wider data set and the process of considering parameters separately for each observation column in the methodology proposed in this study, it seems that our proposed approach is more accurate.
Based on these results, the validation of the methodology is assumed successful. One of the most feature of the proposed methodology in this study is its spatio-temporal flexibility; it is possible to estimate the global atmospheric carbon dioxide mass with acceptable accuracy for desired dates; in addition, since the amount of CO2 mass is calculated for each cell during the procedure, it is possible to define the CO2 mass and its variation in desired locations. A similar approach for CO2 column averaged concentrations was proposed by [89] to investigate the trend over the Indian region. For this purpose, data from SCHIAMACHY and GOSAT were used. Based on the proposed approach, the authors theorize the potential links between seasonal concentrations trends and anthropogeneous behaviour. Similarly, [90] investigated the spatial distribution of the annual average atmospheric CO2 for the state of Mato Grosso (Brazil) using data from OCO2 lite version (V8r). However, both studies do not calculate the CO2 mass amount in the atmosphere as done by the proposed model. Therefore, they do not allow determination of the nominal capacity of the CO2 carbon capture, storage and utilization plants as required for national targets in terms of climate mitigation. In addition, the flexibility provided by the methodology could be useful for control of the anthropogenic activities and to monitor performance of the mitigation and adaptation strategies. However, other aspects such as alteration in carbon sinks should be taken into account since the reported CO2 mass is the balance between carbon source and carbon sink.

4. Conclusions

One of the most important GHGs is carbon dioxide due to its long-lasting presence in the atmosphere and negative impact on climate change. Therefore, the accurate estimation of the atmospheric CO2 mass is crucial to propose mitigation measures and assess their impact. However, the method of calculation has not changed since 1983, even though new measuring platforms, i.e., the satellites, and more data are available. In particular, satellite observation is more reliable for global scale estimation, considering its spatial and temporal resolution. The mean average of the comparison between satellites and TCCON is −0.08 ppm, and the pooled standard deviation is ±1.66 ppm. Among the satellites launched for the purpose of CO2 measurement, the most recently launched satellites, OCO2 and OCO3, were considered to assess their applicability for the new methodology. Due to a low fraction of acceptable observations, after quality flag filtration, and missing days in the OCO3 observation, it was decided to use data from OCO2.
The proposed methodology ensures high resolution to estimate the global atmospheric carbon dioxide using a wide range of observations and better results with respect to the ones that can be obtained by Fraser’s methodology as currently applied. The maximum and average difference between the proposed method and the results of Fraser’s method using NASA data were 1.23% and 0.15%, respectively. Although very accurate models based on observations and chemical transport are available in the literature, the proposed approach could be applied in those cases where computing power or atmospheric data are limited.
Since the proposed methodology divides the Earth into cells according to the satellite spatial resolution, the local and global atmospheric CO2 mass distribution can be assessed. The main issue in this study is related to the availability of data from satellites which might be addressed by combination of satellites or using proper algorithms to reproduce missed data. The result of this study could be helpful in decision-making for the installation of systems for carbon capture and finding the most critical locations in the world to make a plan for climate change mitigation and adaptation. The application of the methodology for such purposes will be shown in a future paper by the authors.

Author Contributions

Conceptualization, M.P., A.A., A.G. and C.S.; Data curation, A.A.; Formal analysis, A.A.; Investigation, A.A.; Methodology, A.A. and A.G.; Project administration, A.G.; Software, A.A.; Supervision, M.P. and A.G.; Validation, M.P., A.G. and C.S.; Visualization, M.P., A.A. and C.S.; Writing—original draft, A.A.; Writing—review and editing, M.P., A.G. and C.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are openly available in www.esrl.noaa.gov and https://disc.gsfc.nasa.gov (accessed on 7 August 2021).

Acknowledgments

The used data were produced by the OCO-2 project at the Jet Propulsion Laboratory, California Institute of Technology, and obtained from the OCO-2 data archive maintained at the NASA Goddard Earth Science Data and Information Services Center.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Figure A1. The number of active observation sites at the beginning of each year. Data elaborated from the information reported in Table A3 in the Appendix A.
Figure A1. The number of active observation sites at the beginning of each year. Data elaborated from the information reported in Table A3 in the Appendix A.
Atmosphere 13 00866 g0a1
Figure A2. Comparison of satellite observations and TCCON stations (References detail are provided in the Table A4 in the Appendix A.). The different studies are labelled based on references.
Figure A2. Comparison of satellite observations and TCCON stations (References detail are provided in the Table A4 in the Appendix A.). The different studies are labelled based on references.
Atmosphere 13 00866 g0a2
Figure A3. OCO2 4-day observation, 22.04.2019–25.04.2019.
Figure A3. OCO2 4-day observation, 22.04.2019–25.04.2019.
Atmosphere 13 00866 g0a3
Figure A4. Global atmospheric carbon dioxide mass 1980–2021.
Figure A4. Global atmospheric carbon dioxide mass 1980–2021.
Atmosphere 13 00866 g0a4
Table A1. Stations data adapted with permission from Fraser et al. [8] for 1980. Copyright 1983 by the American Geophysical Union.flask is sampling and in-situ is the measurement performed in the site.
Table A1. Stations data adapted with permission from Fraser et al. [8] for 1980. Copyright 1983 by the American Geophysical Union.flask is sampling and in-situ is the measurement performed in the site.
No.NameSymbolLatitudeLongitudeFlask (F) or In Situ (I)Concentration a, [ppmv]
1Bass StraitBAS−40°150°F336.7
2Cape GrimCGO−41°145°F335.5
I336.5
3Macquarie IslandMAQ−54°159°F336.9
4MawsonMAW−68°61°F335.5
5Amsterdam IslandAMS−38°78°F337.4
6Ascension IslandASC−8°−14°F338.5
7AzoresAZO38°−27°F340.2
8BarrowBRW71°−157°F340.7
I339.7
9Cape KumukahiKUM20°−145°F340
10Cold BayCBA55°−163°F340.2
11GuamGUA13°145°F340.9
12Key BiscayneKEY26°−80°F340.7
13Mauna LoaMLO20°−156°F341.2
I338.1
14Mould BayMOB76°−119°F340.5
15Niwot RidgeNWR40°−105°F340.4
16Point Six MountPSM47°−114°F341
17SamoaSMO−14°−170°F337.9
I337.9
18SeychellesSEY−5°55°F338.6
19South PoleSPO−90°0F336.1
I335.9
F337
20St. CroixAVI18°−65°F339.9
21Fanning IslandFAN−159°F339.1
Note: a the concentration was referred to 1980.
Table A2. Effective parameters on carbon dioxide concentration.
Table A2. Effective parameters on carbon dioxide concentration.
No.AuthorYear of PublishmentCO2 Sources 1Meteorological Parameters 2Atmospheric Boundary Layer Height Cycle 3Vegetation and ClimatePopulation, GDP and EmploymentSurface Complexity and AlbedoSoil Respiration and Terrestrial EcosystemSampling Time 4Observation Characteristics and HeightWildfirePhytoplankton and OceanAerosols, Clouds and Fog
1M. N. Patil et al. [91]2020
2Yanfen Li et al. [92]2019
3Zhaleh Siabi et al. [93]2019
4Ivakhov V. et al. [94]2019
5Swma Jamalalden Al-jaf, Osama Tareq Al-Taai [95]2019
6Irène Xueref-Remy et al. [96]2018
7Mahesh Patakothi et al. [97]2018
8Shuai Yin et al. [98]2018
9Michael Buchwitz et al. [67]2018
10 Nian Bie et al. [99]2018
11Ge Han et al. [100]2018
12Xun Jiang et al. [101]2017
13Seyed Mohsen Mousavi et al. [102]2017
14Samereh Falahatkar et al. [103]2017
15LEI Li Ping et al. [104]2017
16Debra Wunch et al. [41]2017
17Jacob K. Hedelius et al. [105]2017
18Koorosh Esteki et al. [106]2017
19Chen Pan et al. [107]2016
20Thomas E. Taylor et al. [108]2016
21Yeonjin Jung et al. [109]2016
22Hernández-Paniagua et al. [110]2015
23Min Liu et al. [111]2015
24LIU Xiao-Man et al. [112]2015
25H. Ohyama et al. [52]2015
26Loretta Gratani and Laura Varone [113]2014
27Moon-Soo Park et al. [114]2014
28S. X. Fang et al. [115]2014
29Qin XC et al. [116]2014
30Yanli Li et al. [117]2014
31M. Górka and D. Lewicka-Szczebak [118]2013
32Li Yan-li et al. [119]2013
33Christian Büns and Wilhelm Kuttler [120]2012
34Jiabing Wu et al. [121]2012
35Ma Ángeles García et al. [122]2012
36Yanfang H. [123]2012
37Andrew Rice and Gregory Bostrom [124]2011
38Ramamurthy P. and Pardyjak ER. [125]2011
39Irène Xueref-Remy et al. [126]2011
40Ke Wang et al. [127]2011
41Nawo Eguchi et al. [51]2011
42Y. Yoshida et al. [128]2011
43C. Sirignano et al. [129]2010
44Ch. Gurk et al. [130]2008
45George L. H. Ziska et al. [131]2007
46I. Aben et al. [132]2007
47Yang Y. et al. [133]2006
48Loretta Gratani and Laura Varone [134]2005
49Hassan A. Nasrallah et al. [135]2003
50P. Chamard et al. [136]2003
51Yuesi et al. [137]2002
52Elizabeth A. Wentz et al. [138]2002
53Richard J. Engelen et al. [139]2001
54T. J. Conway et al. [140]1988
1. anthropogenic sources such as urban sources (heating, and traffic), industry; 2. wind speed and direction, precipitation, humidity, temperature, pressure, solar radiation and drought, La Niña and El Niño events; 3. depends on the elevation of the location; 4. in the afternoon for well mixing of the air.
Table A3. Sites that are currently included in the global gas reference network [141].
Table A3. Sites that are currently included in the global gas reference network [141].
SiteNameLocationFirst Carbon Dioxide DatasetStatus Carbon Dioxide DatasetAir Sample Collection Method
1Airborne Aerosol ObservatoryBondville (USA)07.06.2006Terminated 18.09.2009Airborne Flasks *
2Arembepe, BahiaBrazil27.10.2006Terminated 13.01.2010Surface Flasks *
3Alaska Coast GuardUnited States30.04.2009Terminated 21.10.2017Airborne Flasks *
4Alert, NunavutCanada10.06.1985ongoingSurface Flasks
5Amsterdam IslandFrance05.01.1979Terminated 07.12.1990Surface Flasks *
6Argyle, MaineUnited States18.09.2003
22.11.2008
Terminated 29.12.2008
ongoing
In Situ Tall Tower
Surface Flasks
7Anmyeon-doRepublic of Korea03.12.2013ongoingSurface Flasks
8Ascension IslandUnited Kingdom27.08.1979ongoingSurface Flasks
9AssekremAlgeria12.09.1995ongoingSurface Flasks
10St. Croix, Virgin IslandsUnited States16.02.1979Terminated 29.08.1990Surface Flasks *
11Terceira Island, AzoresPortugal26.12.1979ongoingSurface Flasks
12Baltic SeaPoland31.08.1992Terminate 22.06.2011Surface Flasks *
13Boulder Atmospheric Observatory, ColoradoUnited States16.08.2007Terminated 06.07.2016In Situ Tall Tower *
Airborne Flasks *
Surface Flasks *
14Bradgate, IowaUnited States13.09.2004Terminated 18.11.2005Airborne Flasks *
15Baring Head StationNew Zealand14.10.1999ongoingSurface Flasks
16Bukit KototabangIndonesia08.01.2004ongoingSurface Flasks
17St. Davids Head, BermudaUnited Kingdom13.02.1989Terminated 25.01.2010Surface Flasks *
18Tudor Hill, BermudaUnited Kingdom11.05.1989ongoingSurface Flasks
19Beaver Crossing, NebraskaUnited States15.09.2004Terminated 11.05.2011Airborne Flasks *
20Barrow Atmospheric Baseline ObservatoryUnited States25.04.1971ongoingIn Situ Observatory
Surface Flasks
21Black Sea, ConstantaRomania11.10.1994Terminated 26.12.2011Surface Flasks *
22Brentwood, MarylandUnited States25.09.2018ongoingSurface Flasks
23Briggsdale, ColoradoUnited States09.11.1992ongoingAirborne Flasks
24Cold Bay, AlaskaUnited States21.08.1978ongoingSurface Flasks
25Cape Grim, TasmaniaAustralia19.04.1984ongoingSurface Flasks
26Christmas Island Republic of Kiribati08.03.1984ongoingSurface Flasks
27CherskiiRussiaNot for CO2Not for CO2Surface In Situ *
28Centro de Investigacion de la Baja Atmosfera (CIBA)Spain05.05.2009ongoingSurface Flasks
29Offshore Cape May, New JerseyUnited States17.08.2005ongoingAirborne Flasks
30Cape Meares, OregonUnited States10.03.1982Terminate 18.03.1998Surface Flasks *
31CosmosPeru23.06.1979Terminate 28.05.1985Surface Flasks *
32Cape PointSouth Africa11.02.2010ongoingSurface Flasks
33Carbon in Arctic Reservoirs Vulnerability Experiment (CARVE)United States29.06.2012ongoingIn Situ Tall Tower
Airborn Flasks *
Surface Flasks
34Crozet IslandFrance03.03.1991ongoingSurface Flasks
35Dahlen, North DakotaUnited States21.09.2004Terminated 15.11.2016Airborne Flasks
36Drake PassageN/A07.04.2003ongoingSurface Flasks
37Dongsha IslandTaiwan05.03.2010ongoingSurface Flasks
38Easter IslandChile04.01.1994ongoingSurface Flasks
39Estevan Point, British ColumbiaCanada22.11.2002ongoingAirborne Flasks
40East Trout Lake, SaskatchewanCanada15.10.2005Terminated 22.03.2020Airborne Flasks *
41Falkland IslandsUnited Kingdom31.10.1980Terminated 04.02.1982Surface Flasks *
42FortalezaBrazil09.12.2000Terminated 25.03.2003Airborne Flasks *
43Fairchild, WisconsinUnited States20.09.2004Terminated 18.11.2005Airborne Flasks *
44Mariana IslandsGuam24.09.1978ongoingSurface Flasks
45Dwejra Point, GozoMalta11.10.1993Terminated 12.02.1999Surface Flasks *
46Molokai Island, HawaiiUnited States31.05.1999Terminated 22.04.2008Airborne Flasks *
47Halley Station, AntarcticaUnited Kingdom17.01.1983ongoingSurface Flasks
48Harvard Forest, MassachusettsUnited States02.03.2016ongoingAirborne Flasks *
Surface Flasks
49Homer, IllinoisUnited States16.09.2004ongoingAirborne Flasks
50HohenpeissenbergGermany06.04.2006ongoingSurface Flasks
51Humboldt State UniversityUnited States17.05.2008Terminated 31.05.2017Surface Flasks *
52HegyhatsalHungary02.03.1993ongoingSurface Flasks
53Storhofdi, VestmannaeyjarIceland02.10.1992ongoingSurface Flasks
54INFLUX (Indianapolis Flux Experiment)United States09.10.2010ongoingAirborne Flasks *
Surface Flasks
55Grifton, North CarolinaUnited States30.07.1992Terminated 09.06.1999In Situ Tall Tower *
Surface Flasks *
56Izana, Tenerife, Canary IslandsSpain16.11.1991ongoingSurface Flasks
57KaashidhooRepublic of Maldives02.03.1998Terminated 15.07.1999Surface Flasks *
58Key Biscayne, FloridaUnited States13.12.1972ongoingSurface Flasks
59Kitt Peak, ArizonaUnited States20.12.1982Terminated 31.10.1989Surface Flasks *
60Cape Kumukahi, HawaiiUnited States12.01.1971ongoingSurface Flasks
61Sary TaukumKazakhstan12.10.1997Terminated 15.08.2009Surface Flasks *
62Plateau AssyKazakhstan15.10.1997Terminated 05.08.2009Surface Flasks *
63LA MegacitiesUnited States05.11.2014Terminated 08.10.2017Surface Flasks *
64Park Falls, WisconsinUnited States29.11.1994
05.10.2006
ongoing
ongoing
In Situ Tall Tower
Airborne Flasks
Surface Flasks
65Lewisburg, PennsylvaniaUnited States28.02.2013ongoingSurface Flasks
66Lac La Biche, AlbertaCanada30.01.2008Terminated 26.02.2013Surface Flasks *
67LulinTaiwan01.08.2006ongoingSurface Flasks
68LampedusaItaly12.10.2006ongoingSurface Flasks
69Mould Bay, Northwest TerritoriesCanada13.04.1980Terminated 26.05.1997Surface Flasks *
70Mt. Bachelor ObservatoryUnited States14.10.2011
03.05.2012
ongoingongoingSurface Flasks
Surface in situ
71McMurdo Station, AntarcticaUnited States04.12.1985Terminated 28.10.1987Surface Flasks *
72High Altitude Global Climate Observation CenterMexico09.01.2009ongoingSurface Flasks
73Mace Head, County GalwayIreland03.06.1991ongoingSurface Flasks
74Sand Island, MidwayUnited States03.05.1985ongoingSurface Flasks
75Mt. KenyaKenya23.12.2003Terminated 21.06.2011Surface Flasks *
76Mauna Kea, HawaiiUnited StatesN/AN/ASurface Flasks *
77Mauna Loa, HawaiiUnited StatesN/AN/ASurface Flasks *
78Mauna Loa, HawaiiUnited States20.08.1969ongoingIn Situ Observatory
Surface Flasks
79Marcellus PennsylvaniaUnited States03.08.2015ongoingAirborne Flasks *
Surface Flasks
80Mashpee, MassachusettsUnited States11.05.2016ongoingSurface Flasks
81Marthas Vineyard, MassachusettsUnited States27.04.2007Terminated 04.03.2011Surface Flasks *
82Mt. Wilson ObservatoryUnited States30.04.2010ongoingSurface Flasks
83Farol De Mae Luiza LighthouseBrazil12.09.2010Terminated 11.03.2020Surface Flasks *
84NE Baltimore, MarylandUnited States04.04.2018ongoingSurface Flasks
85Offshore Portsmouth, New Hampshire (Isles of Shoals)United States12.09.2003ongoingAirborne Flasks
86GobabebNamibia13.01.1997ongoingSurface Flasks
87NW BaltimoreUnited States17.04.2018ongoingSurface Flasks
88Niwot Ridge Forest, ColoradoUnited States20.01.2006Terminated 08.11.2009Surface Flasks *
89Niwot Ridge, ColoradoUnited States18.05.1967
16.09.2005
ongoing
ongoing
Airborne Flasks *
Surface Flasks
90Kaitorete SpitNew Zealand26.10.1982Terminated 09.04.1985Surface Flasks *
91Oglesby, IllinoisUnited States16.09.2004Terminated 19.11.2005Airborne Flasks *
92Olympic Peninsula, WashingtonUnited States06.01.1984Terminated 30.05.1990Surface Flasks *
93OchsenkopfGermany13.03.2003Terminated 04.06.2019Surface Flasks *
94Pallas-Sammaltunturi, GAW StationFinland21.12.2001ongoingSurface Flasks
95Pico, AzoresPortugal02.08.2010Terminated 18.07.2011Surface Flasks *
96Poker Flat, AlaskaUnited States27.06.1999ongoingAirborne Flasks
97Pacific Ocean (0 N)N/A20.12.1986Terminated 10.07.2017Surface Flasks *
98Pacific Ocean (5 N)N/A19.12.1986Terminated 11.07.2011Surface Flasks *
99Pacific Ocean (10 N)N/A14.01.1987Terminated 12.07.2017Surface Flasks *
100Pacific Ocean (15 N)N/A17.12.1986Terminated 13.07.2017Surface Flasks *
101Pacific Ocean (20 N)N/A16.12.1986Terminated 14.07.2017Surface Flasks *
102Pacific Ocean (25 N)N/A15.12.1986Terminated 15.07.2017Surface Flasks *
103Pacific Ocean (30 N)N/A14.12.1986Terminated 16.07.2017Surface Flasks *
104Pacific Ocean (35 N)N/A21.01.1987Terminated 18.06.2007Surface Flasks *
105Pacific Ocean (40 N)N/A04.06.1987Terminated 14.08.1996Surface Flasks *
106Pacific Ocean (45 N)N/A05.06.1987Terminated 15.08.1996Surface Flasks *
107Pacific Ocean (5 S)N/A21.12.1986Terminated 09.07.2017Surface Flasks *
108Pacific Ocean (10 S)N/A22.12.1986Terminated 08.07.2017Surface Flasks *
109Pacific Ocean (15 S)N/A25.12.1986Terminated 07.07.2017Surface Flasks *
110Pacific Ocean (20 S)N/A28.12.1986Terminated 05.07.2017Surface Flasks *
111Pacific Ocean (25 S)N/A29.12.1986Terminated 04.07.2017Surface Flasks *
112Pacific Ocean (30 S)N/A29.12.1986Terminated 03.07.2017Surface Flasks *
113Pacific Ocean (35 S)N/A30.12.1986Terminated 03.01.2012Surface Flasks *
114Palmer Station, AntarcticaUnited States27.01.1978ongoingSurface Flasks
115Point Six Mountain, MontanaUnited States28.04.1978Terminated 24.12.1982Surface Flasks *
116Point Arena, CaliforniaUnited States05.01.1999Terminated 25.05.2011Surface Flasks *
117Ragged PointBarbados14.11.1987ongoingSurface Flasks
118RarotongaCook Islands16.04.2000ongoingAirborne Flasks
119SantaremBrazil07.12.2000Terminated 20.08.2003Airborne Flasks *
120Offshore Charleston, South CarolinaUnited States22.08.2003ongoingAirborne Flasks
121South China Sea (3 N)N/A05.07.1991Terminated 07.10.1998Surface Flasks *
122South China Sea (6 N)N/A05.07.1991Terminated 09.10.1998Surface Flasks *
123South China Sea (9 N)N/A06.07.1991Terminated 10.10.1998Surface Flasks *
124South China Sea (12 N)N/A06.07.1991Terminated 10.10.1998Surface Flasks *
125South China Sea (15 N)N/A07.07.1991Terminate 15.10.1998Surface Flasks *
126South China Sea (18 N)N/A08.07.1991Terminated 14.10.1998Surface Flasks *
127South China Sea (21 N)N/A08.07.1991Terminated 14.10.1998Surface Flasks *
128Beech Island, South CarolinaUnited States14.08.2008ongoingin Situ Observatory
Surface Flasks
129ShangdianziChina03.09.2009Terminated 02.09.2015Surface Flasks *
130Mahe IslandSeychelles15.01.1980ongoingSurface Flasks
131Bird Island, South GeorgiaUnited Kingdom02.02.1989Terminated 13.08.1992Surface Flasks *
132Southern Great Plains, OklahomaUnited States02.04.2002
29.10.2010
ongoingongoingAirborne Flasks
Surface Flasks
133Shemya Island, AlaskaUnited States04.09.1985ongoingSurface Flasks
134La Jolla, CaliforniaUnited States01.01.1968Terminated 25.09.1986Surface Flasks *
135TutuilaAmerican Samoa15.01.1972ongoingin Situ Observatory
Surface Flasks
136Shenandoah National ParkUnited States26.08.2008ongoingSurface in Situ
137South Pole, AntarcticaUnited States21.01.1975ongoingin Situ Observatory
Surface Flasks
138Ocean Station CharlieUnited States21.11.1968Terminated 12.05.1973Surface Flasks *
139Ocean Station MNorway08.03.1981Terminated 27.11.2009Surface Flasks *
140Sutro Tower, San Francisco, CaliforniaUnited States02.10.2007ongoingSurface Flasks
141SummitGreenland23.06.1997ongoingSurface Flasks
142Syowa Station, AntarcticaJapan25.01.1986ongoingSurface Flasks
143TacolnestonUnited Kingdom06.06.2014Terminated 04.01.2016Surface Flasks *
144Tae-ahn PeninsulaRepublic of Korea24.11.1990ongoingSurface Flasks
145TambopataPeruN/AN/ASurface in Situ
146Offshore Corpus Christi, TexasUnited States09.09.2003ongoingAirborne Flasks
147Trinidad Head, CaliforniaUnited States19.04.2002Terminated 01.06.2017Airborne Flasks
Surface Flasks *
148Hydrometeorological Observatory of TiksiRussia15.08.2011Terminated 03.09.2018Surface Flasks *
149Thurmont, MarylandUnited States01.08.2017ongoingSurface Flasks
150Taiping IslandTaiwan28.05.2019ongoingSurface Flasks
151UlaanbaatarMongolia25.03.2004Terminated 05.03.2009Airborne Flasks *
152UshuaiaArgentina14.09.1994ongoingSurface Flasks
153Wendover, UtahUnited States06.05.1993ongoingSurface Flasks
154Ulaan UulMongolia01.01.1992ongoingSurface Flasks
155West Branch, IowaUnited States28.06.2007ongoingIn Situ Tall Tower
Airborn Flasks
Surface Flasks
156Walnut Grove, CaliforniaUnited States20.09.2007ongoingIn Situ Tall Tower
Airborne Flasks *
Surface Flasks
157Weizmann Institute of Science at the Arava Institute, KeturaIsrael27.11.1995ongoingSurface Flasks
158Moody, TexasUnited States11.02.2001
07.07.2006
Terminated 01.10.2010
ongoing
In Situ Tall Tower
Surface Flasks
159Mt. WaliguanPeoples Republic of China05.08.1990ongoingSurface Flasks
160Western Pacific Cruise (0 N)N/A10.05.2004Terminated 27.05.2013Surface Flasks *
161Western Pacific Cruise (5 N)N/A11.05.2004Terminated 29.05.2013Surface Flasks *
162Western Pacific Cruise (10 N)N/A11.05.2004Terminated 29.05.2013Surface Flasks *
163Western Pacific Cruise (15 N)N/A12.05.2004Terminated 30.05.2013Surface Flasks *
164Western Pacific Cruise (20 N)N/A12.05.2004Terminated 31.05.2013Surface Flasks *
165Western Pacific Cruise (25 N)N/A13.05.2003Terminated 01.06.2013Surface Flasks *
166Western Pacific Cruise (30 N)N/A14.05.2004Terminated 01.06.2013Surface Flasks *
167Western Pacific Cruise (5 S)N/A09.05.2004Terminated 27.05.2013Surface Flasks *
168Western Pacific Cruise (10 S)N/A08.05.2004Terminated 26.05.2013Surface Flasks *
169Western Pacific Cruise (15 S)N/A08.05.2004Terminated 25.05.2013Surface Flasks *
170Western Pacific Cruise (20 S)N/A07.05.2004Terminated 25.05.2013Surface Flasks *
171Western Pacific Cruise (25 S)N/A06.05.2004Terminated 24.05.2013Surface Flasks *
172Western Pacific Cruise (30 S)N/A05.05.2004Terminated 23.05.2013Surface Flasks *
173Ny-Alesund, SvalbardNorway and Sweden11.02.1994ongoingSurface Flasks
*—Indicates discontinued site or project.
Table A4. Comparison of satellite data and TCCON stations.
Table A4. Comparison of satellite data and TCCON stations.
No.LocationCoordinationPeriodSatellite NameSatellite Difference with Ground StationNumber of ObservationsNoteReference
1ChinaNH2010 to 2016GOSAT −1.04 ± 2.10 ppmchinese textcorrelation coefficient of 0.90 Deng A. et al., 2020 [142]
227 TCCON stations July 2009–May 2016GOSAT 0.24 ± 1.68 ppm1913 NH
575 SH
0.349 ± 1.699 ppm NH
−0.128 ± 1.561 ppm SH
2488 matched observationsYawen Kong et al., 2019 [143]
September 2014–July 2017OCO-20.34 ± 1.57 ppm779 NH
294 SH
0.283 ± 1.584 ppm NH
0.494 ± 1.127 ppm SH
1073 matched observations
3Tsukuba36.05° N, 140.12° ESeptember 2014–August 2016 GOSAT0.07± 2.36 ppm N/A Qin et al., 2019 [48]
4Various TCCON stationsBoth HemisphereSeptember 2014 and July 2016OCO-2−0.02 ± 1.36 ppm 34,560RemoTeC algorithm
version 7 data
Lianghai Wu et al., 2018 [144]
5Burgos, Ilocos Norte, Philippines 18.52° N, 120.65° E2017GOSAT−0.86 ± 1.06 ppm N/A Voltaire A. Velazco et al., 2017 [145]
OCO-2−0.83 ± 1.22 ppm 164
6Various TCCON stations OCO-20.4± 1.50 ppm 2790 Wunch et al., 2017 [41]
7global TCCON stations GOSAT0.01 ± 1.22 ppm Zhao-Cheng Zeng et al., 2017 [56]
8Various TCCON stationsBoth Hemisphere2009–2016GOSAT−0.4107 ± 2.216 ppm 1813 NH
596 SH
−0.214 ± 2.009 ppm NH
−1.016 ± 1.956 ppm SH
Ailin Liang et al., 2017 [146]
2014–2016GOSAT−0.62 ± 2.3 ppm 563 NH
151 SH
−0.312 ± 2.006 ppm NH
−1.778 ± 2.096 ppm SH
September 2014 to December 2016OCO-20.2671 ± 1.56 ppm 730 NH
321 SH
0.175 ± 1.402 ppm NH
0.476 ± 1.065 ppm SH
911 TCCON stationsBoth Hemisphere2009–2014GOSAT0.73 ± 1.83 ppm 1484 NH
634 SH
0.959 ± 1.724 ppm NH
0.209 ± 1.706 ppm SH
Photon path length Probability Density Function-Simultaneous (PPDF-S) retrieval methodChisa Iwasaki et al., 2017 [147]
GOSAT−0.32 ± 2.16 ppm 1484 NH
634 SH
−0.299 ± 1.860 ppm NH
−0.384 ± 2.104 ppm SH
standard products for General Users (GU) of XCO2
10Tsukuba and Saga NH GOSAT1.25 ± 2.12 ppm207NIES algorithm Woogyung Kim et al., 2016 [69]
GOSAT1.94 ± 1.89 ppm205ACOS algorithm
11Various TCCON stations September 2014–November 2015OCO-20.87 ± 1.8 ppm not provided in the paper Liang A. et al., 2016 [148]
1211 TCCON stations45° S–80° NJune 2009–April 2014GOSAT±1.7 ppmnot provided in the paperACOS b3.5 Susan Kulawik et al., 2016 [149]
45° S–80° NJanuary 2003–April 2012ENVISAT (SCIAMACHY)±2.1 ppmnot provided in the paperBremen Optimal Estimation DOAS, BESD v2.00.08
1312 TCCON stations 2010 to 2012GOSAT0.21 ± 1.85 ppm 2409 NH
915 SH
0.062 ± 1.815 ppm NH
0.597 ± 1.684 ppm SH
ACOS dataAnjian Deng et al., 2016 [150]
2010 to 2012GOSAT−0.69 ± 2.13 ppm 407 NH
191 SH
−0.679 ± 2.103 ppm NH
−0.720 ± 1.401 ppm SH
NIES data (National Institute for Environmental Studies of Japan)
14Izaña
Ascension Island
Darwin
Reunion Island
Wollongong
28.3° N, 16.5° W
7.9° S, 14.3° W
12.4° S, 130.9° E
20.9° S, 55.5° E
34.4° S, 150.8° E
April 2009–May 2014GOSAT−0.184 ± 0.028 ppm 1137 NH
5877 SH
−0.064 ± 0.032 ppm NH
−0.207 ± 0.027 ppm SH
NIES version 02.21Minqiang Zhou et al., 2016 [151]
April 2009–December 2013GOSAT0.038 ± 0.032 ppm 726 NH
6532 SH
0.057 ± 0.056 ppm NH
0.035 ± 0.028 ppm SH
SRON/KIT algorithm, SRFP v2.3.5
April 2009–June 2014GOSAT−0.006 ± 0.019 ppm 1519 NH
8960 SH
−0.001 ± 0.026 ppm NH
−0.007 ± 0.018 ppm SH
ACOS version 3.5
15TCCON stations GOSAT0.15 ± 1.48 ppmnot provided in the papermodification of the algorithm from Institute of Atmospheric Physics, Chinese Academy of Sciences Dongxu Yang et al., 2015 [70]
16Various TCCON stations 2004–2013GOSAT−0.38 ± 1.992 ppm 5522 NH
1530 SH
−0.364 ± 2.078 ppm NH
−0.439 ± 1.640 ppm SH
Bremen Optimal Estimation DOAS (BESD) algorithmJ. Heymann et al., 2015 [152]
ENVISAT (SCIAMACHY)−0.105 ± 2.017 ppm 32,619 NH
15,336 SH
−0.071 ± 2.097 ppm NH
−0.179 ± 1.836 ppm SH
17Eureka, Park Falls, Lamont, Sodankyla, Bialystok, Orleans and Garmisch NHApril 2010 to March 2012GOSAT−0.94 ± 2.26 ppm 659ACOS ZHANG Miao et al., 2014 [153]
GOSAT−1.49 ± 2.27 ppm 755NIES
ENVISAT−1.52 ± 2.91 ppm 378SCIAMACHY
18Various TCCON stations GOSAT−8.85 ± 4.75 ppm 62SWIR L2 product version 01.xxThe old version with low accuracy and precision (neglected in calculations and figure)I. Morino et al., 2011 [154]
Y. Yoshida et al., 2013 [155]
April 2009 to May 2011GOSAT−1.48 ± 2.09 ppm 567 NH
152 SH
−1.485 ± 1.734 ppm NH
−1.447 ± 2.276 ppm SH
SWIR L2 product version 02.xx
719 observations
19Bialystok, Bremen, Orleans, Park falls, Lamont, Darwin, Wollongong 2009–2011GOSAT −0.20 ± 2.26 ppm 467 NH
110 SH
0.214 ± 2.197 ppm NH
−0.035 ± 2.391 ppm SH
577 observationsA. J. Cogan et al., 2012 [156]
20Bialystok, Orleans, Park Falls, Lamont, Darwin, Wollongong April 2009 and July 2010GOSAT−0.05 ± 0.37%
−0.203 ± 2.654 ppm
759 NH
128 SH
−0.528 ± 2.586 ppm NH
1.721 ± 3.029 ppm SH
TANSO-FTS
887 observation
A. Butz et al., 2011 [157]
NH = Northern Hemisphere, SH = Sothern Hemisphere.
Table A5. worldwide ground-based station.
Table A5. worldwide ground-based station.
No.LocationCoordinationElevation * a.s.l.: above Sea LevelPeriodInstrumentNoteReference
1Bharati, the Indian Antarctic research station 69.24° S, 76.11° E35 m a.s.l. *austral summer (January–February) of 2016Li-Cor CO2/H2O analyzer (model Li-840A) Mahesh Patakothi et al., 2018 [97]
2Bahir Dar and Hawassa 11°36′ N, 37°23′ E
07°15′ N, 38°45′ E
1786–1886 m
1708 m a.s.l.
Aeroqual Series 500 portable gas monitor and YuanTe SKY 2000-M4 handheld multi-gas detector correlation coefficient between instrument was 0.986Oluwasinaayomi Faith Kasim et al., 2018 [158]
3Peterhof station (St. Petersburg, Russia)59.88° N, 29.82° E 2009–2017Fourier transform IR spectrometry (FTIR) using a Bruker 125HRtotal error of 4.18 ± 0.02%, with 0.36 ± 0.06% and 4.16 ± 0.02% for random and systematic errors respectivelyVirolainen Ya. A. 2018 [159]
4Hefei, China31°54′ N, 117°10′ E29 m a.s.l.July 2014–April 2016 Bruker IFS 125HR spectrometer and solar tracker
InGaAs detector from July 2015
similar variation phase and seasonal amplitude with Tsukuba TCCON stationWei Wang et al., 2017 [160]
5Ny-Ålesund78.92° N, 11.92° E 2005–2015Bruker IFS 120HR FTIR spectrometerlower sensitivity in the troposphere in comparison to TCCON (by a factor of 2) Matthias Buschmann et al., 2016 [161]
6Karlsruhe49.094° N, 8.4336° E133 m a.s.l.3 February 2012–22 June 2012EM27 spectrometercommercial low-resolution (0.5 cm−1) (FTS)
agreement with Karlsruhe TCCON station, (0.12 ± 0.08)%
Gisi M. et al., 2012 [162]
7China:
Lin’an, Longfengshan, Shangdianzi, and Waliguan
January 2009 to December 2011cavity ring-down spectroscopy systems (G1301, Picarro Inc.) according to Chen et al., 2010; Crosson, 2008, this type of instrument is suitable for making precise measurementS. X. Fang et al., 2014 [115]
8Kitt Peak, Arizona31.9° N, 111.6° W2070 m a.s.l.1977–1995Fourier transform spectrometer on the McMath telescope.precisions better than 0.5%
similar behavior to the Mauna Loa
Zhonghua Yang et al., 2002 [163]
9Tsukuba, Meteorological Research Institute 36°04′ N, 140°07′ E25 m a.s.l.1986–1996NDIR analyzer (Beckman model 864) from 1986–1992
NDIR analyzer (Beckman model 880) from 1992–1994
Hisayuki Yoshikawa Inoue and Hidekadzu Matsueda 1996 [164]
10Mt. Cimone Station, Italy44°11′ N, 10°42′ E2165 m a.s.l.1979–1992URAS-2T NDIR analyzer, from 1979
ULTRAMAT-5E NDIR from 1988
URAS-3G NDIR (to control)
URAS-2T NDIR precision is ±0.3 p.p.m.v.
ULTRAMAT-5E NDIR precision is ±0.1 p.p.m.v.
V.Cundari et al. 1995 [165]
11Izaña, Tenerife, Canary Islands 28°18′ N, 16°29′ W2367 m a.s.l.1984–1988Siemens Ultramat-3 NADIR the samples were representative of free troposphere in the southern part of the North Atlantic because of the high altitude of the locationBeatriz Navascués et al. 1991 [166]
12Amsterdam island 37°47′ S, 77°31′ E 1980–1989non-dispersive infrared analyzer URAS 2T A. Gaudry et al. 1991 [167]
13
La Jolla, California
Mauna Loa, Hawaii
Cape Kumukahi, Hawaii
Fanning island
and South pole
32.9° N, 117.3° W
19.5° N, 155.6° W
19.5° N, 154.8° W
3.9° N, 159.3° W
90° S, 59° E
March 1977–February 1982non-dispersive infrared gas analyzer Illem g. Mook and Marjan Koopmans 1983 [168]
14Shetland Isles, Scotland 60.2° N, 1.2° W 1992–1996Carle Series 400 gas chromatograph
Finnigan MAT 252 mass spectrometer with MT Box-C gas preparationsystem
a part of CSIRO networkR.J. Francey et al. 1998 [169]
15Schauinsland station, southwest Germany47°55’ N, 7°55’ E1205 m a.s.l.1972–2002nondispersive infrared analysis (NDIR)
Until August 1980,URAS-2 (Hartmann & Braun),
from September 1980 until the end of 1993,Ultramat-3 (Siemens)
and from 1994 onward with URAS-3 (Hartmann & Braun)
The accuracy of the data was estimated: better than ±1 ppm for the period 1972–
1991 and better than ±0.5 ppm later on.
M. Schmidt et al., 2003 [170]
16Kasprowy Wierch
Kraków
in southern Poland
49°14′ N, 19°59′ E
50°04′ N, 19°55′ E
1989 m a.s.l.
220 m a.s.l.
1996–2006Automated gas chromatographs (Hewlett Packard, Series 5890, with FID detector and Ni catalyst for conversion of CO2 to CH4 and Porapak Q column) L. Chmura et al., 2008 [171]
17Moscow to Khabarovsk 1997–2004 LI6262 gas analyzer (LICOR, United States) mobile measurement at surface layer with the error of ±1 ppm at a CO2 concentration of 350 ppm.
The intrinsic noise was 0.2 ppm
I. B. Belikov et al., 2006 [172]
18ZOTTO international observatory, Krasnoyarsk krai, Russia60° N, 90° E114 m a.s.l.January 2006-December 2013NDIR CO2 Analyzer (Siemens AG, Ultramat 6F) up to April 2007
EnviroSense 3000i gas-analyzing system (Picarro Inc., USA) from May 2009
using the tall tower (302 m)
measurement error does not exeed 0.1 ppm
A. V. Timokhina et al., 2015 [173]
E. A. Kozlova and A. C. Manning 2009 [174]
19Cabauw51.971° N, 4.927° E −0.7 m a.s.l.1992–2010Siemens Ultramat NDIR 1992–2004
NDIR (LICOR 7000) after 2004
sampling in tall tower
Siemens Ultramat NDIR resolution in the range of 0–500 ppm was 0.5 ppm
A. T. Vermeulen et al., 2011 [175]
20Barrow (Alaska)71.32° N, 156.61° W11.00 m a.s.l.stablished in 1973non-dispersive infrared analyzer https://www.esrl.noaa.gov/gmd/obop/brw/ (accessed on 16 August 2021)
21American Samoa14.24° S, 170.56° W42.00 m a.s.l.stablished in 1974non-dispersive infrared analyzer https://www.esrl.noaa.gov/gmd/obop/smo/ (accessed on 16 August 2021)
https://cdiac.ess-dive.lbl.gov/trends/co2/sio-sam.html (accessed on 16 August 2021)
22South Pole 90° S, 59° E2837 m a.s.l.stablished in 1957non-dispersive infrared gas analyzer https://www.esrl.noaa.gov/gmd/obop/spo/ (accessed on 16 August 2021)
Illem g. Mook and Marjan Koopmans 1983 [168]
23Ascension Island (SH)7.92° S, 14.33° W10 m a.s.l.Available data from 22.05.2012–31.10.2018 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
24Anmeyondo (KR)36.54° N, 126.33° E30 m a.s.l.Available data from 02.02.2015–18.04.2018 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
25Bialystok (PL)53.23° N, 23.025° E180 m a.s.l.Available data from 01.03.2009–01.10.2018 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
26Bremen (DE)53.10° N, 8.85° E27 m a.s.l.Available data from 22.01.2010–23.08.2019 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
27Burgos18.533° N, 120.650° E35 m a.s.l.Available data from 03.03.2017–31.01.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
28Caltech (US)34.136° N, 118.127° W230 m a.s.l.Available data from 20.09.2012–03.10.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
29Darwin (AU)12.42° S, 130.89° E
12.46° S, 130.93° E
30 m a.s.l.
37 m a.s.l.
Available data from 28.08.2005–31.01.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
30Edwards (US) Available data from 20.07.2013–03.10.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
31East Trout Lake54.35° N, 104.99° W501.8 m a.s.l.Available data from 07.10.2016–06.09.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
32Eureka (CA)80.05° N, 86.42° W610 m a.s.l.Available data from 24.07.2010–06.07.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
33Four Corners (US)36.80° N, 108.48° W1643 m a.s.l.Available data from 16.03.2013–04.10.2013 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
34Garmisch (DE)47.476° N, 11.063° E740 m a.s.l.Available data from 16.07.2007–18.10.2019 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
35Hefei (PRC)31.90° N, 118.67° E29 m a.s.l.Available data from 18.09.2015–31.12.2016 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
36Indianapolis (US)39.86° N, 86.00° W270 m a.s.l.Available data from 23.08.2012–01.12.2012 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
37Izana (ES)28.3° N, 16.5° W2370 m a.s.l.Available data from 18.05.2007–02.11.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
38Jet Propulsion Laboratory (US)34.20° N, 118.175° W390 m a.s.l.Available data from 31.07.2007–22.06.2008 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
39Jet Propulsion Laboratory (US)34.20° N, 118.175° W390 m a.s.l.Available data from 19.05.2011–14.05.2018 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
40Saga (JP)33.24° N, 130.29° E7 m a.s.l.Available data from 28.07.2011–04.08.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
41Karlsruhe (DE)49.10° N, 8.44° E116 m a.s.l.Available data from 19.04.2010–31.10.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
42Lauder (NZ)45.04° S, 169.68° E370 m a.s.l.Available data from 29.06.2004–09.12.2010 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
43Lauder (NZ)45.04° S, 169.68° E370 m a.s.l.Available data from 02.02.2010–31.10.2018 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
44Lauder (NZ)45.04° S, 169.68° E370 m a.s.l.Available data from 03.10.2018–31.07.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
45Manaus (BR)3.21° S, 60.59° W50 m a.s.l.Available data from 01.10.2014–24.06.2015 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
46Nicosia35.14° N, 33.38° E185 m a.s.l.Available data from 31.08.2019–31.01.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
47Lamont (US)36.60° N, 97.48° W320 m a.s.l.Available data from 06.07.2008–03.10.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
48Orléans (FR)47.97° N, 2.11° E130 m a.s.l.Available data from 29.08.2009–18.09.2019 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
49Park Falls (US)45.94° N, 90.27° W440 m a.s.l.Available data from 02.06.2004–03.10.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
50Paris (FR)48.84° N, 2.35° E60 m a.s.l.Available data from 23.09.2014–24.01.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
51Réunion Island (RE)20.90° S, 55.48° E87 m a.s.l.Available data from 16.09.2011–18.07.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
52Rikubetsu (JP)43.45° N, 143.77° E380 m a.s.l.Available data from 16.11.2013–30.09.2019 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
53Sodankylä (FI)67.37° N, 26.63° E188 m a.s.l.Available data from 16.05.2009–30.09.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
54Ny Ålesund78.9° N, 11.9° E20 m a.s.l.Available data from 06.04.2014–15.09.2019 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
55Tsukuba (JP)36.05° N, 140.12° E30 m a.s.l.Available data from 04.08.2011–30.09.2019 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
56Wollongong (AU)34.41° S, 150.88° E30 m a.s.l.Available data from 26.06.2008–31.01.2020 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
57Zugspitze (DE)47.42° N, 10.98° E2960 m a.s.l.Available data from 24.04.2015–17.10.2019 TCCON stationhttps://tccondata.org/ (accessed on 16 August 2021)
https://tccon-wiki.caltech.edu/Main/TCCONSites (accessed on 16 August 2021)
Table A6. Tall towers in the NOAA Earth System Research Laboratory’s Global Greenhouse Gas Reference Network [176].
Table A6. Tall towers in the NOAA Earth System Research Laboratory’s Global Greenhouse Gas Reference Network [176].
No.NameLocationCoordinationSurface ElevationIntake HeightCarbon Dioxide Measurement PeriodNote
1Argyle, Maine Tower (AMT)Argyle, Maine45.03° N, 68.68° W50 m a.s.l.12, 30, 107 m above ground2003-ongoinssg
2Boulder Atmospheric Observatory (BAO)Erie, Colorado40.05° N, 105.01° W1584 m a.s.l.22, 100, 300 m above ground2007–2016Discontinued
3Barrow Observatory (BRW)Barrow, Alaska71.323° N, 156.6114° W11 m a.s.l.16.46 m above ground1971-ongoing
4WITN Tower (ITN) Grifton, North Carolina5.53° N, 77.38° W9 m a.s.l.51, 123, 496 m above ground1992–1999Discontinued
5WLEF Tower (LEF)Park Falls, Wisconsin45.9451° N, 90.2732° W472 m a.s.l.1, 30, 76, 122, 244, 396 m above ground2003-ongoing
6Mount Bachelor Observatory (MBO)Mount Bachelor, Oregon43.9775° N, 121.6861° W2731 m a.s.l.11 m above ground2011-ongoing
2012-ongoing
7Mauna Loa Observatory (MLO)Mauna Loa, Hawaii19.5362° N, 155.5763° W3397 m a.s.l.40 m above ground1969-ongoing
8South Carolina Tower (SCT)Beech Island, South Carolina33.406° N, 81.833° W115 m a.s.l.30, 61, 305 m above ground2008-ongoing
9American Samoa Observatory (SMO)Tutuila Island, American Samoa14.2474° S, 170.5644° W42 m a.s.l.18 m above ground1972-ongoing
10Shenandoah National Park (SNP)Shenandoah National Park, Virginia38.617° N, 78.35° W1008 m a.s.l.5, 10, 17 m above ground2008-ongoing
11South Pole Observatory (SPO)South Pole, Antarctica89.98° S, 24.8° W2810 m a.s.l.11 m above ground1975-ongoing
12West Branch, Iowa (WBI)West Branch, Iowa41.725° N, 91.353° W242 m a.s.l.31, 99, 379 m above ground2007-ongoing
13Walnut Grove, California (WGC)Walnut Grove, California38.265° N, 121.4911° W0 m a.s.l.30, 91, 483 m above ground2007-ongoing
14WKT Tower (WKT)Moody, Texas31.32° N, 97.33° W251 m a.s.l.30, 122, 457 m above ground2003-ongoing

References

  1. Brovkin, V.; Lorenz, S.; Raddatz, T.; Ilyina, T.; Stemmler, I. What was the source of the atmospheric CO2 increase during the Holocene? Biogeosciences 2019, 16, 2543–2555. [Google Scholar] [CrossRef] [Green Version]
  2. Pearman, G.I.; Etheridge, D.; Silva, F.d.; Fraser, P.J. Evidence of changing concentrations of atmospheric CO2, N2O and CH4 from air bubbles in Antarctic ice. Nature 1986, 320, 248–250. [Google Scholar] [CrossRef]
  3. Yue, T.; Zhang, L.; Zhao, M.; Wang, Y.; Wilson, J. Space-and ground-based CO2 measurements: A review. Sci. China Earth Sci. 2016, 59, 2089–2097. [Google Scholar] [CrossRef]
  4. Pieter Tans, and Ralph Keeling, Global Monitoring Laboratory, NOAA/GML. Available online: www.esrl.noaa.gov/gmd/ccgg/trends/ (accessed on 10 February 2021).
  5. McElwain, J.; Montañez, I.; White, J.; Wilson, J.; Yiotis, C. Was atmospheric CO2 capped at 1000 ppm over the past 300 million years? Palaeogeogr. Palaeoclimatol. Palaeoecol. 2016, 441, 653–658. [Google Scholar] [CrossRef] [Green Version]
  6. Steinthorsdottir, M.; Vajda, V. Early Jurassic (late Pliensbachian) CO2 concentrations based on stomatal analysis of fossil conifer leaves from eastern Australia. Gondwana Res. 2015, 27, 932–939. [Google Scholar] [CrossRef] [Green Version]
  7. Goldberg, F. Rate of Increasing Concentrations of Atmospheric Carbon Dioxide Controlled by Natural Temperature Variations. Energy Environ. 2008, 19, 995–1011. [Google Scholar] [CrossRef] [Green Version]
  8. Fraser, P.J.; Pearman, G.I.; Hyson, P. The Global Distribution of Atmospheric Carbon Dioxide 2. A Review of Provisional Background Observations, 1978–1980. J. Geophys. Res. 1983, 88, 3591–3598. [Google Scholar] [CrossRef]
  9. Friedlingstein, P.; Jones, M.W.; O’Sullivan, M.; Andrew, R.M.; Hauck, J.; Peters, G.P.; Peters, W.; Pongratz, J.; Sitch, S.; Quéré, C.L.; et al. Global Carbon Budget 2019. Earth Syst. Sci. Data 2019, 11, 1783–1838. [Google Scholar] [CrossRef] [Green Version]
  10. Jiang, X.; Yung, Y.L. Global Patterns of Carbon Dioxide Variability from Satellite Observations. Annu. Rev. Earth Planet. Sci. 2019, 47, 225–245. [Google Scholar] [CrossRef]
  11. NOAA. Carbon Cycle Greenhouse Gases. Available online: https://www.esrl.noaa.gov/gmd/ccgg/ (accessed on 10 February 2021).
  12. TCCON. Available online: https://tccondata.org/ (accessed on 22 May 2020).
  13. CONTRAIL. Available online: https://www.cger.nies.go.jp/contrail/index.html (accessed on 10 February 2021).
  14. Singh, H.B.; Jacob, D.; Pfister, L.; Hipskind, R.S. INTEX-NA: Intercontinental Chemical Transport Experiment—North America; NASA Ames Research Center: Moffett Field, CA, USA, 2001. [Google Scholar]
  15. NCAR Earth Observing Laboratory. HIAPER Pole-to-Pole Observation. Available online: https://www.eol.ucar.edu/field_projects/hippo-1 (accessed on 10 February 2021).
  16. LAGOS. In-Service Aircraft for a Global Observing System. Available online: https://www.iagos.org/ (accessed on 10 February 2021).
  17. NASA. Carbon in Arctic Reservoirs Vulnerability Experiment (CARVE). Available online: https://carve.ornl.gov/ (accessed on 10 February 2021).
  18. NCEI (NOAA National Centers for Environmental Information). Ships Of Opportunity Program (SOOP) Data. Available online: https://www.ncei.noaa.gov/access/ocean-carbon-data-system/oceans/VOS_Program/ (accessed on 17 February 2021).
  19. Takahashi, T.; Sutherland, S.C.; Kozyr, A.; NCEI (NOAA National Centers for Environmental Information). Global Surface pCO2 (LDEO) Database V2019. 2020. Available online: https://www.ncei.noaa.gov/access/ocean-carbon-data-system/oceans/LDEO_Underway_Database/ (accessed on 17 February 2021).
  20. Sutton, A.; Sabine, C.; Mathis, J.; Kozyr, A.; Carbon Dioxide Information Analysis Center, Oak Ridge National Laboratory, US Department of Energy. Available online: https://www.ncei.noaa.gov/access/ocean-carbon-data-system/oceans/Moorings/ndp092.html (accessed on 10 February 2021).
  21. Pan, G.; Xu, Y.; Ma, J. The potential of CO2 satellite monitoring for climate governance: A review. J. Environ. Manag. 2021, 277, 111423. [Google Scholar] [CrossRef]
  22. NASA. Orbiting Carbon Observatory-3. Available online: https://ocov3.jpl.nasa.gov/ (accessed on 10 February 2021).
  23. NASA. Orbiting Carbon Observatory-2. Available online: https://earth.esa.int/eogateway/missions/gosat (accessed on 10 February 2021).
  24. European Space Agency. GOSAT. Available online: https://earth.esa.int/eogateway/missions/gosat (accessed on 10 February 2021).
  25. NASA. TES Get Data. Available online: https://tes.jpl.nasa.gov/data/ (accessed on 10 February 2021).
  26. NASA. AIRS Atmospheric Infrared Sounder. Available online: https://airs.jpl.nasa.gov/data/get-data/standard-data/ (accessed on 10 February 2021).
  27. NOAA. NOAA Comprehensive Large Array-Data Stewardship System (Class). Available online: https://www.avl.class.noaa.gov/saa/products/search?datatype_family=IASI (accessed on 10 February 2021).
  28. eoPortal Directory. SciSat-1/ACE (Science Satellite/Atmospheric Chemistry Experiment). Available online: https://directory.eoportal.org/web/eoportal/satellite-missions/s/scisat-1 (accessed on 10 February 2021).
  29. SCHIAMACHY. Available online: https://www.sciamachy.org/ (accessed on 10 February 2021).
  30. NOAA. ESRL Global Monitoring Division—Mauna Loa Observatory. Available online: https://www.esrl.noaa.gov/gmd/obop/mlo/ (accessed on 10 February 2021).
  31. Wunch, D.; Toon, G.C.; Blavier, J.-F.L.; Washenfelder, R.A.; Notholt, J.; Connor, B.J.; Griffith, D.W.T.; Sherlock, V.; Wennberg, P.O. The total carbon column observing network. Philos. Trans. A Math Phys. Eng. Sci. 2011, 369, 2087–2112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Bakwin, P.S.; Tans, P.P.; Zhao, C.; Ussler, W., III; Quesnell, E. Measurements of carbon dioxide on a very tall tower. Tellus 1995, 47B, 535–549. [Google Scholar] [CrossRef]
  33. Bakwin, P.S.; Tans, P.P.; Hurst, D.F.; Zhao, C. Measurements of carbon dioxide on very tall towers: Results of the NOAA/CMDL program. Tellus B 1998, 50, 401–415. [Google Scholar] [CrossRef] [Green Version]
  34. NOAA. Measuring & Analyzing Greenhouse Gases: Behind the Scenes. Available online: https://www.esrl.noaa.gov/gmd/ccgg/behind_the_scenes/towers.html (accessed on 10 February 2021).
  35. Machida, T.; Matsueda, H.; Sawa, Y.; Nakagawa, Y.; Hirotani, K.; Kondo, N.; Goto, K.; Nakazawa, T.; Ishikawa, K.; Ogawa, T. Worldwide Measurements of Atmospheric CO2 and Other Trace Gas Species Using Commercial Airlines. J. Atmos. Ocean. Technol. 2008, 25, 1744–1754. [Google Scholar] [CrossRef]
  36. Avissar, R.; Holder, H.E.; Abehserra, N.; Bolch, M.A.; Novick, K. The Duke University Helicopter Observation Platform. Bull. Amer. Meteor. Soc. 2009, 90, 939–954. [Google Scholar] [CrossRef] [Green Version]
  37. Ouchi, M.; Masumi, Y.; Nakayama, T.; Shimizu, K.; Sawada, T. Development of a balloon-borne instrument for CO2 vertical profile observations in the troposphere. Atmos. Meas. Tech. 2019, 12, 5639–5653. [Google Scholar] [CrossRef] [Green Version]
  38. Bischof, W. Carbon Dioxide Measurements from Aircraft. Tellus 1970, 22, 545–549. [Google Scholar] [CrossRef] [Green Version]
  39. Hou, Y.; Wang, S.; Zhou, Y.; Yan, F.; Zhu, J. Analysis of the carbon dioxide concentration in the lowest atmospheric layers and the factors affecting China based on satellite observations. Int. J. Remote Sens. 2012, 34, 1981–1994. [Google Scholar] [CrossRef]
  40. Ota, Y.; Imasu, R. CO2 Retrieval Using Thermal Infrared Radiation Observation by Interferometric Monitor for Greenhouse Gases (IMG) Onboard Advanced Earth Observing Satellite (ADEOS). J. Meteorol. Soc. Jpn. 2016, 471-490. [Google Scholar] [CrossRef] [Green Version]
  41. Wunch, D.; Wennberg, P.O.; Osterman, G.; Fisher, B.; Naylor, B. Comparisons of the Orbiting Carbon Observatory-2 (OCO-2) XCO2 measurements with TCCON. Atmos. Meas. Tech. 2017, 10, 2209–2238. [Google Scholar] [CrossRef] [Green Version]
  42. Bi, Y.; Wang, Q.; Yang, Z.; Chen, J.; Bai, W. Validation of Column-Averaged Dry-Air Mole Fraction of CO2 Retrieved from OCO-2 Using Ground-Based FTS Measurements. J. Meteorol. Res. 2018, 32, 433–443. [Google Scholar] [CrossRef]
  43. Timofeyev, Y.M.; Berezin, I.A.; Virolainen, Y.A.; Makarova, M.V.; Polyakov, A.V.; Poberovsk, A.V. Spatial–Temporal CO2 Variations near St. Petersburg Based on Satellite and Ground-Based Measurements. Atmos. Ocean. Phys. 2019, 55, 59–64. [Google Scholar] [CrossRef]
  44. Wu, L.; Hasekamp, O.; Hu, H.; De Brugh, J.A.; Landgraf, J.; Butz, A.; Aben, I. Full-physics carbon dioxide retrievals from the Orbiting Carbon Observatory-2 (OCO-2) satellite by only using the 2.06 µm band. Atmos. Meas. Tech. 2019, 12, 6049–6058. [Google Scholar] [CrossRef] [Green Version]
  45. O’Dell, C.W.; Eldering, A.; Wennberg, P.O.; Crisp, D.; Gunson, M.R.; Fisher, B.; Frankenberg, C.; Kiel, M.; Lindqvist, H.; Mandrake, L.; et al. Improved retrievals of carbon dioxide from Orbiting Carbon Observatory-2 with the version 8 ACOS algorithm. Atmos. Meas. Tech. 2018, 11, 6539–6576. [Google Scholar] [CrossRef] [Green Version]
  46. Ailin, L.; Ge, H.; Wei, G.; Jie, Y.; Chengzhi, X. Comparison of Global XCO2 Concentrations From OCO-2 With TCCON Data in Terms of Latitude Zones. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2017, 10, 2491–2498. [Google Scholar]
  47. Liang, A.; Han, G.; Xu, H.; Gong, W.; Zhang, T. Evaluation of XCO2 from OCO-2 Lite File Product compared with TCCON data. Int. Geosci. Remote Sens. Symp. IGARSS 2017, 8127389, 2070–2073. [Google Scholar]
  48. Qin, X.C.; Nakayama, T.; Matsumi, Y.; Kawasaki, M.; Imasu, R.; Morino, I.; Tanaka, Y.; Ishidoya, S.; Sato, K.; Ohashi, M. Observation of column-averaged molar mixing ratios of carbon dioxide in Tokyo. Atmos. Environ. X 2019, 2, 100022. [Google Scholar]
  49. Dan-dan, L.I.U.; Yin-bo, H.U.A.N.G.; Zhen-song, C.A.O.; Xing-ji, L.U.; Yu-song, S.U.N.; Qian-si, T.U. Analysis of Total Columns of Greenhouse Gas Based on Direct Observation and Comparison with Satellite Data in Hefei. Acta Photonica Sin. 2020, 49, 0301002. [Google Scholar] [CrossRef]
  50. Velazco, V.A.; Deutscher, N.M.; Morino, I.; Uchino, O.; Bukosa, B. Satellite and ground-based measurements of XCO2 in a remote semiarid region of Australia. Earth Syst. Sci. Data 2019, 11, 935–946. [Google Scholar] [CrossRef] [Green Version]
  51. Eguchi, N.; Yoshida, Y.; Morino, I.; Kikuchi, N.; Saeki, T. Seasonal variations of greenhouse gas column-averaged dry air mole fractions retrieved from swir spectra of gosat tanso-fts. In Proceedings of the 2011 IEEE International Geoscience and Remote Sensing Symposium (IGRASS), Vancouver, BC, Canada, 24–29 July 2011; pp. 3542–3545. [Google Scholar] [CrossRef]
  52. Ohyama, H.; Kawakami, S.; Tanaka, T.; Morino, I.; Uchino, O. Observations of XCO2 and XCH4 with ground-based high-resolution FTS at Saga, Japan, and comparisons with GOSAT products. Atmos. Meas. Tech. 2015, 8, 5263–5276. [Google Scholar] [CrossRef] [Green Version]
  53. Rokotyan, N.V.; Imasu, R.; Zakharov, V.I.; Gribanov, K.G.; Khamatnurova, M.Y. The Amplitude of the CO2 Seasonal Cycle in the Atmosphere of the Ural Region Retrieved from Ground—Based and Satellite Near—IR Measurements. Atmos. Ocean. Opt. 2015, 28, 49–55. [Google Scholar] [CrossRef]
  54. Yates, E.L.; Schiro, K.; Lowenstein, M.; Sheffner, E.J.; Iraci, L.T. Carbon Dioxide and Methane at a Desert Site—A Case Study at Railroad Valley Playa, Nevada, USA. Atmosphere 2011, 2, 702–714. [Google Scholar] [CrossRef] [Green Version]
  55. Qu, Y.; Zhang, C.; Wang, D.; Tian, P.; Bai, W.; Zhang, X. Comparison of atmospheric CO2 observed by GOSAT and two ground stations in China. Int. J. Remote Sens. 2014, 34, 3938–3946. [Google Scholar] [CrossRef]
  56. Zeng, Z.-C.; Lei, L.; Strong, K.; Jones, D.B.; Guo, L. 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]
  57. Wunch, D.; Wennberg, P.O.; Toon, G.C.; Connor, B.J.; Fisher, B.; Osterman, G.B.; Frankenberg, C. A Method for Evaluating Bias in Global Measurements of CO2 Total Columns from Space. Atmos. Chem. Phys. 2011, 23, 12317–12337. [Google Scholar] [CrossRef] [Green Version]
  58. Yuan, Y.; Sussmann, R.; Rettinger, M.; Ries, L.; Petermeier, H.; Menzel, A. Comparison of Continuous In-Situ CO2 Measurements with Co-Located Column-Averaged XCO2 TCCON/Satellite Observations and CarbonTracker Model Over the Zugspitze Region. Remote Sens. 2019, 11, 2981. [Google Scholar] [CrossRef] [Green Version]
  59. Buchwitz, M.; Reuter, M.; Schneising, O.; Hewson, W.; Detmers, R. Global satellite observations of column-averaged carbon dioxide and methane: The GHG-CCI XCO2 and XCH4 CRDP3 data set. Remote Sens. Environ. 2017, 203, 276–295. [Google Scholar] [CrossRef] [Green Version]
  60. Buchwitz, M.; Reuter, M.; Schneising, O.; Boesch, H.; Guerlet, S.; Dils, B.; Aben, I.; Armante, R.; Bergamaschi, P.; Blumenstock, T.; et al. The Greenhouse Gas Climate Change Initiative (GHG-CCI): Comparison and quality assessment of near-surface-sensitive satellite-derived CO2 and CH4 global data sets. Remote Sens. Environ. 2013, 162, 344–362. [Google Scholar] [CrossRef] [Green Version]
  61. Miao, R.; Lu, N.; Yao, L.; Zhu, Y.; Wang, J.; Sun, J. Multi-Year Comparison of Carbon Dioxide from Satellite Data with Ground-Based FTS Measurements (2003–2011). Remote Sens. 2013, 5, 3431–3456. [Google Scholar] [CrossRef] [Green Version]
  62. Moon, Y.S.; Arellano, A.F. The characteristics of tropospheric CO2 retrieved by AIRS, GOSAT and IASI in East Asia. Disaster Adv. 2015, 8, 1–13. [Google Scholar]
  63. Zhang, L.; Zhang, X.; Jiang, H. Accuracy comparisons of airs, sciamachy and gosat with ground-based data based on global CO2 concentration. In Proceedings of the 21st International Conference on Geoinformatics, Kaifen, China, 20–22 June 2013; pp. 1–5. [Google Scholar]
  64. Zhang, L.; Jiang, H.; Zhang, X. Comparison analysis of the global carbon dioxide concentration column derived from sciamachy, airs, and gosat with surface station measurements. Int. J. Remote Sens. 2015, 36, 1406–1423. [Google Scholar] [CrossRef]
  65. Jiang, X.; Crisp, D.; Olsen, E.T.; Kulawik, S.S.; Miller, C.E. CO2 annual and semiannual cycles from multiple satellite retrievals and models. Earth Space Sci. 2016, 3, 78–87. [Google Scholar] [CrossRef]
  66. Reuter, M.; Buchwitz, M.; Schneising, O.; Noël, S.; Bovensmann, H.; Burrows, J.P.; Boesch, H.; Di Noia, A.; Anand, J.; Parker, R.J.; et al. Ensemble-based satellite-derived carbon dioxide and methane column-averaged dry-air mole fraction data sets (2003–2018) for carbon and climate applications. Atmos. Meas. Tech. 2020, 13, 789–819. [Google Scholar] [CrossRef] [Green Version]
  67. Buchwitz, M.; Reuter, M.; Schneising, O.; Noëll, S.; Gier, B. Computation and analysis of atmospheric carbon dioxide annual mean growth rates from satellite observations during 2003–2016. Atmos. Chem. Phys. 2018, 18, 17355–17370. [Google Scholar] [CrossRef] [Green Version]
  68. Dils, B.; Buchwitz, M.; Reuter, M.; Schneising, O.; Boesch, H.; Parker, R.; Guerlet, S.; Aben, I.; Blumenstock, T.; Burrows, J.P.; et al. The Greenhouse Gas Climate Change Initiative (GHG-CCI): Comparative validation of GHG-CCI sciamachy/envisat and tanso-fts/gosat CO2 and CH4 retrieval algorithm products with measurements from the TCCON. Atmos. Meas. Tech. 2014, 7, 1723–1744. [Google Scholar] [CrossRef] [Green Version]
  69. Kim, W.; Kim, J.; Jung, Y.; Boesch, H.; Lee, H. Retrieving XCO2 from GOSAT FTS over East Asia Using Simultaneous Aerosol Information from CAI. Remote Sens. 2016, 8, 994. [Google Scholar] [CrossRef] [Green Version]
  70. Yang, D.; Liu, Y.; Cai, Z.; Deng, J.; Wang, J.; Chen, X. An advanced carbon dioxide retrieval algorithm for satellite measurements and its application to GOSAT observations. Sci. Bull. 2015, 60, 2063–2066. [Google Scholar] [CrossRef] [Green Version]
  71. Lindqvist, H.; O’Dell, C.W.; Basu, S.; Boesch, H.; Chevallier, F.; Deutscher, N.; Feng, L.; Fisher, B.; Hase, F.; Inoue, M.; et al. Does GOSAT capture the true seasonal cycle of carbon dioxide? Atmos. Chem. Phys. 2015, 15, 13023–13040. [Google Scholar] [CrossRef] [Green Version]
  72. Yi, Y.; Liu, Y.; Cai, Z.; Fang, S.; Yang, D.; Wang, Y.; Liang, M.; Yan, B.; Ma, Q.; Wang, M. Measuring and comparing in-situ CO2 and CO profiles with satellite observations and model data. Atmos. Ocean. Sci. Lett. 2019, 12, 444–450. [Google Scholar] [CrossRef] [Green Version]
  73. Tadić, J.M.; Michalak, A.M. On the effect of spatial variability and support on validation of remote sensing observations of CO2. Atmos. Environ. 2016, 132, 309–316. [Google Scholar] [CrossRef] [Green Version]
  74. Maddy, E.S.; Barnet, C.D.; Goldberg, M.; Sweeney, C.; Liu, X. CO2 retrievals from the Atmospheric Infrared Sounder: Methodology and validation. J. Geophys. Res. Atmos. 2008, 113, D11301. [Google Scholar] [CrossRef] [Green Version]
  75. Chahine, M.; Barnet, C.; Olsen, E.T.; Chen, L.; Maddy, E. On the determination of atmospheric minor gases by the method of vanishing partial derivatives with application to CO2. Geophys. Res. Lett. 2005, 32. [Google Scholar] [CrossRef] [Green Version]
  76. Uspensky, A.B.; Kukharsky, A.V.; Romanov, S.V.; Rublev, A.N. Monitoring the Carbon Dioxide Mixing Ratio in the Troposphere and the Methane Total Column over Siberia According to the Data of the AIRS and IASI IR Sounders. Atmos. Ocean. Phys. 2011, 47, 1097–1103. [Google Scholar] [CrossRef]
  77. Kukharskii, A.V.; Uspenskii, A.B. Determination of tropospheric mean carbon dioxide concentration from satellite high spectral resolution IR-sounder data. Russ. Meteorol. Hydrol. 2009, 34, 202–211. [Google Scholar] [CrossRef]
  78. Frankenberg, C.; Kulawik, S.S.; Wofsy, S.C.; Chevallier, F.; Daube, B.; Kort, E.A.; O’dell, C.; Olsen, E.T.; Osterman, G. Using airborne HIAPER Pole-to-Pole Observations (HIPPO) to evaluate model and remote sensing estimates of atmospheric carbon dioxide. Atmos. Chem. Phys. 2016, 16, 7867–7878. [Google Scholar] [CrossRef] [Green Version]
  79. Glen, S. StatisticsHowTo.com. Available online: https://www.statisticshowto.com/pooled-standard-deviation/ (accessed on 15 January 2021).
  80. NASA. GES DISK. Available online: https://disc.gsfc.nasa.gov/datasets/OCO2_L2_Lite_FP_9r/summary (accessed on 10 February 2021).
  81. NASA. User’s Guide. Available online: https://docserver.gesdisc.eosdis.nasa.gov/public/project/OCO/OCO2_DUG.V9.pdf (accessed on 12 November 2021).
  82. Yu, S.; Rosenberg, R.; Bruegge, C.; Chapsky, L.; Fu, D.; Lee, R.; Taylor, T.; Cronk, H.; O’Dell, C.; Angal, A.; et al. Stability Assessment of OCO-2 Radiometric Calibration Using Aqua MODIS as a Reference. Remote Sens. 2020, 12, 1269. [Google Scholar] [CrossRef] [Green Version]
  83. NASA. Available online: https://ocov2.jpl.nasa.gov/observatory/instrument/ (accessed on 11 November 2021).
  84. NASA. GES DISK Documentation. Available online: https://docserver.gesdisc.eosdis.nasa.gov/public/project/OCO/OCO2_OCO3_B10_DUG.pdf (accessed on 10 February 2021).
  85. Pearman, G.I.; Hyson, P.; Fraser, P.J. The global distribution of atmospheric carbon dioxide: 1 Aspects of observations and modeling. J. Geophys. Res. 1983, 88, 3581–3590. [Google Scholar] [CrossRef]
  86. Trenberth, K.E.; Guillemot, C.J. The total mass of the atmosphere. J. Geophys. Res. 1994, 99, 23079–23088. [Google Scholar] [CrossRef]
  87. Trenberth, K.E.; Smith, L. The Mass of the Atmosphere: A Constraint on Global Analyses. J. Clim. 2005, 18, 864–875. [Google Scholar] [CrossRef]
  88. Müller, A.; Tanimoto, H.; Sugita, T.; Machida, T.; Nakaoka, S.-I.; Patra, P.K.; Laughner, J.; Crisp, D. New approach to evaluate satellite derived XCO2 over oceans by integrating ship and aircraft observations. Atmos. Chem. Phys. 2021, 21, 8255–8271. [Google Scholar] [CrossRef]
  89. Prasad, P.; Rastogi, S.; Singh, R. Study of satellite retrieved CO2 and CH4 concentration over India. Adv. Space Res. 2014, 54, 1933–1940. [Google Scholar] [CrossRef]
  90. Rossi, F.S.; Santos, G.A.D.A.; Maria, L.D.S.; Lourençoni, T.; Pelissari, T.D.; Della-Silva, J.L.; Júnior, J.W.O.; Silva, A.d.A.e.; Lima, M.; Teodoro, P.E.; et al. Carbon dioxide spatial variability and dynamics for contrasting land uses in central Brazil agricultural frontier from remote sensing data. J. S. Am. Earth Sci. 2022, 116, 103809. [Google Scholar] [CrossRef]
  91. Patil, M.N.; Dharmaraj, T.; Waghmare, R.T.; Singh, S.; Pithani, P.; Kulkarni, R.; Dhangar, N.; Siingh, D.; Chinthalu, G.R.; Singh, R.; et al. Observations of carbon dioxide and turbulent fluxes during fog conditions in north India. J. Earth Syst. Sci. 2020, 129, 51. [Google Scholar] [CrossRef]
  92. Li, Y.; Zhang, C.; Li, F. Influence of thin cirrus clouds on the short-wave infrared satellite observation of atmospheric CO2. Opt. Int. J. Light Electron Opt. 2019, 208, 164140. [Google Scholar] [CrossRef]
  93. Siabi, Z.; Falahatkar, S.; Alavi, S.J. Spatial distribution of XCO2 using OCO-2 data in growing seasons. J. Environ. Manag. 2019, 244, 110–118. [Google Scholar] [CrossRef]
  94. Ivakhov, V.M.; Paramonova, N.N.; Privalov, V.I.; Zinchenko, A.V.; Loskutova, M.A.; Makshtas, A.P.; Kustov, V.Y.; Laurila, T.; Aurela, M.; Asmi, E. Atmospheric Concentration of Carbon Dioxide at Tiksi and Cape Baranov Stations in 2010–2017. Russ. Meteorol. Hydrol. 2019, 44, 291–299. [Google Scholar] [CrossRef]
  95. Al-jaf, S.J.; Al-Taai, O.T. Impact of carbon dioxide concentrations on atmospheric temperature changes over iraq and some neighboring countries. Plant Arch. 2019, 19, 1450–1456. [Google Scholar]
  96. Xueref-Remy, I.; Dieudonné, E.; Vuillemin, C.; Lopez, M.; Lac, C.; Schmidt, M.; Delmotte, M.; Chevallier, F.; Ravetta, F.; Perrussel, O.; et al. Diurnal, synoptic and seasonal variability of atmospheric CO2 in the Paris megacity area. Atmos. Chem. Phys. 2018, 18, 3335–3362. [Google Scholar] [CrossRef] [Green Version]
  97. Pathakoti, M.; Gaddamidi, S.; Gharai, B.; Syamala, P.S.; Rao, P.V. Influence of meteorological parameters on atmospheric CO2 at Bharati, the Indian Antarctic research station. Polar Res. 2018, 37, 1442072. [Google Scholar] [CrossRef] [Green Version]
  98. Yin, S.; Wang, X.; Tani, H.; Zhang, X.; Zhong, G. Analyzing temporo-spatial changes and the distribution of the CO2 T concentration in Australia from 2009 to 2016 by greenhouse gas monitoring satellites. Atmos. Environ. 2018, 192, 1–12. [Google Scholar] [CrossRef] [Green Version]
  99. Bie, N.; Lei, L.; Zeng, Z.; Cai, B.; Yang, S. Regional uncertainty of GOSAT XCO2 retrievals in China: Quantification and attribution. Atmos. Meas. Tech. 2018, 11, 1251–1272. [Google Scholar] [CrossRef] [Green Version]
  100. Ge, H.; Hao, X.; Wei, G.; Jiqiao, L.; Juan, D. Feasibility Study on Measuring Atmospheric CO2 in Urban Areas Using Spaceborne CO2-IPDA LIDAR. Remote Sensig 2018, 10, 985. [Google Scholar]
  101. Jiang, X.; Kao, A.; Corbett, A.; Olsen, E.; Pagano, T.; Zhai, A.; Newman, S.; Li, L.; Yung, Y. Influence of Droughts on Mid-Tropospheric CO2. Remote Sens. 2017, 9, 852. [Google Scholar] [CrossRef] [Green Version]
  102. Mousavi, S.M.; Falahatkar, S.; Farajzadeh, M. Assessment of seasonal variations of carbon dioxide concentration in Iran using GOSAT data. Nat. Resour. Forum 2017, 41, 83–91. [Google Scholar] [CrossRef]
  103. Falahatkar, S.; Mousavi, S.M.; Farajzadeh, M. Spatial and temporal distribution of carbon dioxide gas using GOSAT data over IRAN. Environ. Monit. Assess. 2017, 189, 627. [Google Scholar] [CrossRef] [PubMed]
  104. Lei, L.; Zhong, H.; He, Z.; Cai, B.; Yang, S.; Wu, C.; Zeng, Z.; Liu, L.; Zhang, B. Assessment of atmospheric CO2 concentration enhancement from anthropogenic emissions based on satellite observations. Chin. J. 2017, 62, 2941–2950. [Google Scholar]
  105. Hedelius, J.K.; Feng, S.; Roehl, C.M.; Wunch, D.; Hillyard, P.W. Emissions and topographic effects on column CO2 (X CO2) variations, with a focus on the Southern California Megacity. J. Geophys. Res. Atmos. 2017, 122, 7200–7215. [Google Scholar] [CrossRef]
  106. Esteki, K.; Prakash, N.; Li, Y.; Mu, C.; Du, K. Seasonal Variation of CO2 Vertical Distribution in the Atmospheric Boundary Layer and Impact of Meteorological Parameters. Int. J. Environ. Res. 2017, 11, 707–721. [Google Scholar] [CrossRef]
  107. Pan, C.; Zhu, X.; Wei, N.; Zhu, X.; She, Q. Spatial variability of daytime CO2 concentration with landscape structure across urbanization gradients, Shanghai, China. Clim. Res. 2016, 69, 107–116. [Google Scholar] [CrossRef]
  108. Taylor, T.E.; O’Dell, C.W.; Frankenberg, C.; Partain, P.T.; Cronk, H.Q. Orbiting Carbon Observatory-2 (OCO-2) cloud screening algorithms: Validation against collocated MODIS and CALIOP data. Atmos. Meas. Tech. 2016, 9, 973–989. [Google Scholar] [CrossRef] [Green Version]
  109. Jung, Y.; Kim, J.; Kim, W.; Lee, H.B. Impact of Aerosol Property on the Accuracy of a CO2 Retrieval Algorithm from Satellite Remote Sensing. Remote Sens. 2016, 8, 322. [Google Scholar] [CrossRef] [Green Version]
  110. Hernández-Paniagua, I.Y.; Lowry, D.; Clemitshaw, K.C.; Fisher, R.E.; France, J.L.; Lanoisellé, M.; Ramonet, M.; Nisbet, E.G. Diurnal, seasonal, and annual trends in atmospheric CO2 at southwest London during 2000–2012: Wind sector analysis and comparison with Mace Head, Ireland. Atmos. Environ. 2015, 105, 138–147. [Google Scholar] [CrossRef]
  111. Liu, M.; Wu, J.; Zhu, X.; He, H.; Jia, W.; Xiang, W. Evolution and variation of atmospheric carbon dioxide concentration over terrestrial ecosystems as derived from eddy covariance measurements. Atmos. Environ. 2015, 114, 75–82. [Google Scholar] [CrossRef] [Green Version]
  112. Liu, X.M.; Cheng, X.L.; Hu, F. Gradient characteristics of CO2 concentration and flux in Beijing urban area part I: Concentration and virtual temperature. Chin. J. Geophys. Chin. Ed. 2015, 17, 501–511. [Google Scholar]
  113. Gratani, L.; Varone, L. Atmospheric carbon dioxide concentration variations in Rome: Relationship with traffic level and urban park size. Urban Ecosyst. 2014, 17, 501–511. [Google Scholar] [CrossRef]
  114. Park, M.-S.; Joo, S.J.; Park, S.-U. Carbon dioxide concentration and flux in an urban residential area in Seoul, Korea. Adv. Atmos. Sci. 2014, 31, 1101–1112. [Google Scholar] [CrossRef]
  115. Fang, S.X.; Zhou, L.X.; Tans, P.P.; Ciais, P.; Steinbacher, M.; Xu, L.; Luan, T. In situ measurement of atmospheric CO2 at the four WMO/GAW stations in China. Atmos. Chem. Phys. 2014, 14, 2541–2554. [Google Scholar] [CrossRef] [Green Version]
  116. Retrieval and analysis of atmospheric XCO2 using ground-based spectral observation. Guang Pu Xue Yu Guang Pu Fen Xi/Spectrosc. Spectr. Anal. 2014, 34, 1729–1735.
  117. Li, Y.; Deng, J.; Mu, C.; Xing, Z.; Du, K. Vertical distribution of CO2 in the atmospheric boundary layer: Characteristics and impact of meteorological variables. Atmos. Environ. 2014, 91, 110–117. [Google Scholar] [CrossRef]
  118. Górka, M.; Lewicka-Szczebak, D. One-year spatial and temporal monitoring of concentration and carbon isotopic composition of atmospheric CO2 in a Wrocław (SW Poland) city area. Appl. Geochem. 2013, 35, 7–13. [Google Scholar] [CrossRef]
  119. Li, Y.L.; Mu, C.; Deng, J.J.; Zhao, S.H.; Du, K. Near surface atmospheric CO2 variations in autumn at suburban Xiamen, China. Huanjing Kexue/Environ. Sci. 2013, 34, 2018–2024. [Google Scholar]
  120. Büns, C.; Kuttler, W. Path-integrated measurements of carbon dioxide in the urban canopy layer. Atmos. Environ. 2012, 237–247. [Google Scholar] [CrossRef]
  121. Wu, J.; Guan, D.; Yuan, F.; Yang, H.; Wang, A.; Jin, C. Evolution of atmospheric carbon dioxide concentration at different temporal scales recorded in a tall forest. Atmos. Environ. 2012, 61, 9–14. [Google Scholar] [CrossRef]
  122. García, M.A.; Sánchez, M.L.; Pérez, I.A. Differences between carbon dioxide levels over suburban and rural sites in Northern Spain. Environ. Sci. Pollut. Res. 2012, 19, 432–439. [Google Scholar] [CrossRef]
  123. Yanfang, H.; Shixin, W.; Yi, Z.; Feng, W.; Wenliang, L.; Yu, H. Variations of Greenhouse Gas carbon dioxide concentration over the Chinese mainland based on Satellite and Background Measurements. Disaster Adv. 2012, 5, 871–876. [Google Scholar]
  124. Bostrom, A.R. Measurements of carbon dioxide in an Oregon metropolitan region. Atmos. Environ. 2011, 45, 1138–1144. [Google Scholar]
  125. Ramamurthy, P.; Pardyjak, E.R. Toward understanding the behaviour of carbon dioxide and surface energy fluxes in the urbanized semi-arid Salt Lake Valley, Utah, USA. Atmos Environ. 2011, 45, 73–84. [Google Scholar] [CrossRef]
  126. Xueref-Remy, I.; Messager, C.; Filippi, D.; Pastel, M.; Nedelec, P.; Ramonet, M.; Paris, J.D.; Ciais, P. Variability and budget of CO2 in Europe: Analysis of the CAATER airborne campaigns—Part 1: Observed variability. Atmos. Chem. Phys. 2011, 11, 5655–5672. [Google Scholar] [CrossRef] [Green Version]
  127. Wang, K.; Jiang, H.; Zhang, X.; Zhou, G. Analysis of spatial and temporal variations of carbon dioxide over China using SCIAMACHY satellite observations during 2003–2005. Int. J. Remote Sens. 2011, 32, 815–832. [Google Scholar] [CrossRef]
  128. Yoshida, Y.; Ota, Y.; Eguchi, N.; Kikuchi, N.; Nobuta, K.; Tran, H.; Morino, I.; Yokota, T. Retrieval algorithm for CO2 and CH4 column abundances from short-wavelength infrared spectral observations by the Greenhouse gases observing satellite. Atmos. Meas. Tech. 2011, 4, 717–734. [Google Scholar] [CrossRef] [Green Version]
  129. Sirignano, C.; Neubert, R.E.; Rödenbeck, C.; Meijer, H.A. Atmospheric oxygen and carbon dioxide observations from two European coastal stations 2000–2005: Continental influence, trend changes and APO climatology. Atmos. Chem. Phys. 2010, 10, 1599–1615. [Google Scholar] [CrossRef] [Green Version]
  130. Gurk, C.; Fischer, H.; Hoor, P.; Lawrence, M.G.; Lelieveld, J.; Wernli, H. Airborne in-situ measurements of vertical, seasonal and latitudinal distributions of carbon dioxide over Europe. Atmos. Chem. Phys. 2008, 8, 6395–6403. [Google Scholar] [CrossRef] [Green Version]
  131. George, K.; Ziska, L.H.; Bunce, J.A.; Quebedeauxb, B. Elevated atmospheric CO2 concentration and temperature across an urban–rural transect. Atmos. Environ. 2007, 41, 7654–7665. [Google Scholar] [CrossRef] [Green Version]
  132. Aben, I.; Hasekamp, O.; Hartmann, W. Uncertainties in the space-based measurements of CO2 columns due to scattering in the Earth’s atmosphere. J. Quant. Spectrosc. Radiat. Transf. 2007, 104, 450–459. [Google Scholar] [CrossRef]
  133. Yang, Y.; Tan, J.; Zheng, Y.; Cheng, S. Study on the atmospheric stabilities and the thickness of atmospheric mixed layer during recent 15 years in Shanghai. Sci. Meteorol. Sin. 2006, 26, 536–541. [Google Scholar]
  134. Gratani, L.; Varone, L. Daily and seasonal variation of CO2 in the city of Rome in relationship with the traffic volume. Atmos. Environ. 2005, 39, 2619–2624. [Google Scholar] [CrossRef]
  135. Nasrallah, H.A.; Balling, R.C., Jr.; Madi, S.M.; Al-Ansari, L. Temporal variations in atmospheric CO2 concentrations in Kuwait City, Kuwait with comparisons to Phoenix, Arizona, USA. Environ. Pollut. 2003, 121, 301–305. [Google Scholar] [CrossRef]
  136. Chamard, P.; Thiery, F.; Sarra, A.D.; Ciattaglia, L.; Silvestri, L.D. Interannual variability of atmospheric CO2 in the Mediterranean: Measurements at the island of Lampedusa. Tellus 2003, 55B, 83–93. [Google Scholar] [CrossRef]
  137. Yuesi, W.; Changke, W.; Xueqing, G.; Guangren, L.; Yao, H. Trend, seasonal and diurnal variations of atmospheric CO2 in Beijing. Chin. Sci. Bull. 2002, 47, 2050–2055. [Google Scholar]
  138. Wentz, E.A.; Gober, P.; Balling, R.C., Jr.; Day, T.A. Spatial Patterns and Determinants of Winter Atmospheric Carbon Dioxide Concentrations in an Urban Environment. Ann. Assoc. Am. Geogr. 2002, 92, 15–28. [Google Scholar] [CrossRef]
  139. Engelen, R.J.; Denning, A.S.; Gurney, K.R.; Stephen, G.L. Global observations of the carbon budget 1 Expected satellite capabilities for emission spectroscopy in the EOS and NPOESS eras. J. Geophys. Res. 2001, 106, 20055–20068. [Google Scholar] [CrossRef] [Green Version]
  140. Conway, T.J.; Tans, P.; Waterman, L.S.; Thoning, K.W.; Masarie, K.A.; Gammon, R.H. Atmospheric carbon dioxide measurements in the remote global troposphere, 1981–1984. Tellus B Chem. Phys. Meteorol. 1988, 40, 81–115. [Google Scholar] [CrossRef]
  141. NOAA. Global Monitoring Laboratory. Available online: https://www.esrl.noaa.gov/gmd/dv/site/index.php?program=ccgg (accessed on 10 February 2021).
  142. Deng, A.J.; Guo, H.B.; Hu, J.; Jiang, C.Z.; Liu, P.G.; Jing, H.F. Temporal and distribution characteristic of CO2 concentration over China based on GOSAT satellite data. J. Remote Sens. 2020, 24, 319–325. [Google Scholar]
  143. Kong, Y.; Chen, B.; Measho, S. Spatio-Temporal Consistency Evaluation of XCO2 Retrievals from GOSAT and OCO-2 Based on TCCON and Model Data for Joint Utilization in Carbon Cycle Research. Atmosphere 2019, 10, 354. [Google Scholar] [CrossRef] [Green Version]
  144. Wu, L.; Hasekamp, O.; Hu, H.; Landgraf, J.; Butz, A. Carbon dioxide retrieval from OCO-2 satellite observations using the RemoTeC algorithm and validation with TCCON measurements. Atmos. Meas. Tech. 2018, 11, 3111–3130. [Google Scholar] [CrossRef] [Green Version]
  145. Velazco, V.A.; Morino, I.; Uchino, O.; Hori, A.; Kie, M. TCCON Philippines: First Measurement Results, Satellite Data and Model Comparisons in Southeast. Asia Remote Sens. 2017, 9, 1228. [Google Scholar] [CrossRef] [Green Version]
  146. Ailin, L.; Wei, G.; Ge, H.; Chengzhi, X. Comparison of Satellite-Observed XCO2 from GOSAT, OCO-2, and Ground-Based TCCON. Remote Sens. 2017, 9, 1033. [Google Scholar]
  147. Chisa, I.; Ryoichi, I.; Andrey, B.; Tatsuya, Y.; Yukio, Y. Validation of GOSAT SWIR XCO2 and XCH4 Retrieved by PPDF-S Method and Comparison with Full Physics Method. Sci. Online Lett. Atmos. 2017, 13, 168–173. [Google Scholar]
  148. Liang, A.; Gong, W.; Han, G. OCO-2 XCO2 validation using TCCON data. Int. Geosci. Remote Sens. Symp. IGARSS 2016, 7729210, 830–833. [Google Scholar]
  149. Kulawik, S.; Wunch, D.; O’Dell, C.; Frankenberg, C.; Reuter, M. Consistent evaluation of ACOS-GOSAT, BESD-SCIAMACHY, CarbonTracker, and MACC through comparisons to TCCON. Atmos. Meas. Tech. 2016, 9, 683–709. [Google Scholar] [CrossRef] [Green Version]
  150. Anjian, D.; Tao, Y.; Tianhai, C.; Xingfa, G.; Fengjie, Z.; Hong, G. Intercomparison of Carbon Dioxide Products Retrieved from GOSAT Short-Wavelength Infrared Spectra for Three Years (2010–2012). Atmosphere 2016, 7, 109. [Google Scholar]
  151. Zhou, M.; Dils, B.; Wang, P.; Detmers, R.; Yoshida, Y. Validation of TANSO-FTS/GOSAT XCO2 and XCH4 glint mode retrievals using TCCON data from near-ocean sites. Atmos. Meas. Tech. 2016, 9, 1415–1430. [Google Scholar] [CrossRef] [Green Version]
  152. Heymann, J.; Reuter, M.; Hilker, M.; Buchwitz, M.; Schneising, O. Consistent satellite XCO2 retrievals from SCIAMACHY and GOSAT using the BESD algorithm. Atmos. Meas. Tech. 2015, 8, 2961–2980. [Google Scholar] [CrossRef] [Green Version]
  153. Miao, Z.; Xing-Ying, Z.; Rui-Xia, L.; Lie-Qun, H. A study of the validation of atmospheric CO2 from satellite hyper spectral remote sensing. Adv. Clim. Chang. Res. 2014, 5, 131–135. [Google Scholar]
  154. Morino, I.; Uchino, O.; Inoue, M.; Yoshida, Y.; Yokota, T. Preliminary validation of column-averaged volume mixing ratios of carbon dioxide and methane retrieved from GOSAT short-wavelength infrared spectra. Atmos. Meas. Tech. 2011, 4, 1061–1076. [Google Scholar] [CrossRef] [Green Version]
  155. Yoshida, Y.; Kikuchi, N.; Morino, I.; Uchino, O.; Oshchepkov, S. Improvement of the retrieval algorithm for GOSAT SWIR XCO2 and XCH4 and their validation using TCCON data. Atmos. Meas. Tech. 2013, 6, 1533–1547. [Google Scholar] [CrossRef]
  156. Cogan, A.J.; Boesch, H.; Parker, R.J.; Feng, L.; Palmer, P.I.; Blavier, J.F.; Deutscher, N.M.; Macatangay, R.; Notholt, J.; Roehl, C.; et al. Atmospheric carbon dioxide retrieved from the Greenhouse gases Observing SATellite (GOSAT): Comparison with ground-based TCCON observations and GEOS-Chem model calculations. J. Geophys. Res. Atmos. 2012, 117, D21. [Google Scholar] [CrossRef] [Green Version]
  157. Butz, A.; Guerlet, S.; Hasekamp, O.; Schepers, D.; Galli, A.; Aben, I.; Frankenberg, C.; Hartmann, J.M.; Tran, H.; Kuze, A.; et al. Toward accurate CO2 and CH4 observations from GOSAT. Geophys. Res. Lett. 2011, 38, L14812. [Google Scholar] [CrossRef] [Green Version]
  158. Kasim, O.F.; Abshare, M.W.; Mukuna, T.E.; Wahab, B. Land Use and Ambient Air Quality in Bahir Dar and Hawassa, Ethiopia. Air Soil Water Res. 2018, 11, 1–10. [Google Scholar] [CrossRef]
  159. Virolainen, Y.A. Methodical aspects of the determination of carbon dioxide in atmosphere using ftir spectroscopy. J. Appl. Spectrosc. 2018, 85, 462–469. [Google Scholar] [CrossRef]
  160. Wang, W.; Tian, Y.; Liu, C.; Sun, Y.; Liu, W. Investigating the performance of a greenhouse gas observatory in Hefei, China. Atmos. Meas. Tech. 2017, 10, 2627–2643. [Google Scholar] [CrossRef] [Green Version]
  161. Buschmann, M.; Deutscher, N.M.; Sherlock, V.; Palm, M.; Warneke, T.; Notholt, J. Retrieval of xCO2 from ground-based mid-infrared (NDACC) solar absorption spectra and comparison to TCCON. Atmos. Meas. Tech. 2016, 9, 577–585. [Google Scholar] [CrossRef] [Green Version]
  162. Gisi, M.; Hase, F.; Dohe, S.; Blumenstock, T.; Simon, A.; Keens, A. XCO2-measurements with a tabletop FTS using solar absorption spectroscopy. Atmos. Meas. Tech. 2012, 5, 2969–2980. [Google Scholar] [CrossRef] [Green Version]
  163. Yang, Z.; Toon, G.C.; Margolis, J.S.; Wennberg, P.O. Atmospheric CO2 retrieved from ground-based near IR solar spectra. Geophys. Res. Lett. 2002, 29, 53-1–53-4. [Google Scholar] [CrossRef] [Green Version]
  164. Matsueda, I.H. Variations in Atmospheric CO2 at the Meteorological Research Institute, Tsukuba, Japan. J. Atmos. Chem. 1996, 23, 137–161. [Google Scholar]
  165. Cundari, V.; Colombo, T.; Ciattagli, L. Thirteen Years of Atmospheric Carbon Dioxide Measurements at Mt Cimone Station 1995, Italy. Il Nuovo Cim. 1995, 18C, 33–47. [Google Scholar] [CrossRef]
  166. Navascues, B.; Jardin, C.; Rus, C. Carbon dioxide observations at Izaña baseline station, Tenerife (Canary Islands): 1984–1988. Tellus B 1991, 43, 118–125. [Google Scholar] [CrossRef] [Green Version]
  167. Gaudry, A.; Monfray, P.; Polian, G.; Bonsang, G.; Ardouin, B.; Jegou, A.; Lambert, G. Non-seasonal variations of atmospheric CO2 concentrations at Amsterdam Island. Tellus B Chem. Phys. Meteorol. B 1991, 43, 136–143. [Google Scholar] [CrossRef]
  168. Mook, W.G.; Koopmans, M.; Carter, A.F.; Keeling, C.D. Seasonal, Latitudinal, and Secular Variations in the Abundance and Isotopic Ratios of Atmospheric Carbon Dioxide 1 Results From Land Stations. J. Geophys. Res. 1983, 88, 10915–10933. [Google Scholar] [CrossRef]
  169. Francey, R.J.; Steele, L.P.; Langenfelds, R.L.; Allison, C.E.; Cooper, L.N. Atmospheric carbon dioxide and its stable isotope ratios, methane, carbon monoxide, nitrous oxide and hydrogen from Shetland Isles. Atmos. Environ. 1998, 32, 3331–3338. [Google Scholar] [CrossRef]
  170. Schmidt, M.; Graul, R.; Sartorius, H.; Levin, I. The Schauinsland CO2 record: 30 years of continental observations and their implications for the variability of the European CO2 budget. J. Geophys. Res. Atmos. 2003, 108. [Google Scholar] [CrossRef]
  171. Chmura, L.; Rozanski, K.; Necki, J.M.; Zimnoch, M.; Korus, A.; Pycia, M. Atmospheric Concentrations of Carbon Dioxide in Southern Poland: Comparison of Mountain and Urban Environments. Pol. J. Environ. Stud. 2008, 6, 859–867. [Google Scholar]
  172. Belikov, I.B.; Brenninkmeijer, C.A.; Elansky, N.F.; Ral’ko, A.A. Methane, Carbon Monoxide, and Carbon Dioxide Concentrations Measured in the Atmospheric Surface Layer over Continental Russia in the TROICA Experiments. Atmos. Ocean. Phys. 2006, 42, 46–59. [Google Scholar] [CrossRef]
  173. Timokhina, A.V.; Prokushkin, A.S.; Onuchin, A.A.; Panov, A.V.; Kofman, G.B. Long-term Trend in CO2 Concentration in the Surface Atmosphere over Central Siberia. Russ. Meteorol. Hydrol. 2015, 40, 186–190. [Google Scholar] [CrossRef]
  174. Kozlova, E.A. Methodology and calibration for continuous measurements of biogeochemical trace gas and O2 concentrations from a 300-m tall tower in central Siberia. Atmos. Meas. Tech. 2009, 2, 205–220. [Google Scholar] [CrossRef] [Green Version]
  175. Vermeulen, A.T.; Hensen, A.; Popa, M.E.; Bulk, W.C.M.V.D.; Jongejan, P.A.C. Greenhouse gas observations from Cabauw Tall Tower (1992–2010). Atmos. Meas. Tech. 2011, 4, 617–644. [Google Scholar] [CrossRef] [Green Version]
  176. NOAA. In-Situ Measurement Program. Available online: https://gml.noaa.gov/ccgg/insitu/ (accessed on 10 February 2021).
Figure 1. Building blocks for the global atmospheric carbon dioxide content calculation. Letters (ag) are used to label the blocks in order to simplify mentioning the steps in the text.
Figure 1. Building blocks for the global atmospheric carbon dioxide content calculation. Letters (ag) are used to label the blocks in order to simplify mentioning the steps in the text.
Atmosphere 13 00866 g001
Figure 2. Building blocks for the proposed methodology to reach a higher accuracy of the results for global atmospheric carbon content. Letters (aj) are used to label the blocks in order to simplify mentioning the steps in the text.
Figure 2. Building blocks for the proposed methodology to reach a higher accuracy of the results for global atmospheric carbon content. Letters (aj) are used to label the blocks in order to simplify mentioning the steps in the text.
Atmosphere 13 00866 g002
Figure 3. OCO2 global 4-day carbon dioxide mass estimation in the period 2019–2021. The orange is 4-day global atmospheric carbon dioxide mass estimation, the black curve is related to NOAA data, and the blue is the fitting curve. The red dashed line is the prediction of global carbon dioxide mass since the data was not available for OCO2.
Figure 3. OCO2 global 4-day carbon dioxide mass estimation in the period 2019–2021. The orange is 4-day global atmospheric carbon dioxide mass estimation, the black curve is related to NOAA data, and the blue is the fitting curve. The red dashed line is the prediction of global carbon dioxide mass since the data was not available for OCO2.
Atmosphere 13 00866 g003
Table 1. KPI table of the means of observations.
Table 1. KPI table of the means of observations.
Means of ObservationAccuracy (a)PrecisionCost (b)CoverageTime Scale
Ground-based stationsCould be ±0.5 ppmBetter than 0.25% for TCCONlow-high (c)local-regional (d)Sampling and Analysing duration
Aircraft, Helicopters and BalloonsCould be less than ±0.25 ppm (±0.1, ±0.05 ppm are also obtained)Could be ±0.1 ppmlow-high (c)regionalFlight and Analysing duration (e)
ShipsN/A (a)Could be better than 0.6%lowregionalTravel duration + Analysing duration (c)
Satellites−0.08 ppm regarding TCCON according to the calculationsLess than 2 ppm is neededhighglobalOCO2 & OCO3-16 days GOSAT-3 days
Notes: (a) It depends on the utilized instrument. (b) The satellite cost is assumed to be the comparison base. (c) In case of network it is expensive. (d) Regarding the elevation of the sampling vessel and filtering approach, this can be representative of local or regional. In case of combination of the stations in a network with models it is possible to have the global coverage. (e) It can be only sampling and analyzing duration if the analyzing instruments are installed in the means of observation.
Table 2. Comparison of the global atmospheric carbon dioxide mass using NOAA data and OCO2 data, 2019–2020.
Table 2. Comparison of the global atmospheric carbon dioxide mass using NOAA data and OCO2 data, 2019–2020.
NOAA CO2 Mass ResultsOCO2 Data CO2 Mass Results
Average value (kg)3.201 × 10153.206 × 1015
Maximum value (kg)3.226 × 10153.248 × 1015
Minimum value (kg)3.167 × 10153.167 × 1015
Maximum difference ((OCO2–NOAA)/NOAA)1.23%
Average difference0.15%
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pellegrini, M.; Aghakhani, A.; Guzzini, A.; Saccani, C. Modification of Fraser’s Method for the Atmospheric CO2 Mass Estimation by Using Satellite Data. Atmosphere 2022, 13, 866. https://doi.org/10.3390/atmos13060866

AMA Style

Pellegrini M, Aghakhani A, Guzzini A, Saccani C. Modification of Fraser’s Method for the Atmospheric CO2 Mass Estimation by Using Satellite Data. Atmosphere. 2022; 13(6):866. https://doi.org/10.3390/atmos13060866

Chicago/Turabian Style

Pellegrini, Marco, Arash Aghakhani, Alessandro Guzzini, and Cesare Saccani. 2022. "Modification of Fraser’s Method for the Atmospheric CO2 Mass Estimation by Using Satellite Data" Atmosphere 13, no. 6: 866. https://doi.org/10.3390/atmos13060866

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