Next Article in Journal
Attribution Analysis of Runoff Change in the Upper Reaches of the Kaidu River Basin Based on a Modified Budyko Framework
Previous Article in Journal
Spatiotemporal Analysis of PM2.5 Concentrations on the Incidence of Childhood Asthma in Developing Countries: Case Study of Cartagena de Indias, Colombia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Retrieved XCO2 Accuracy Improvement by Reducing Aerosol-Induced Bias for China’s Future High-Precision Greenhouse Gases Monitoring Satellite Mission

1
State Key Laboratory of Modern Optical Instrumentation, College of Optical Science and Engineering, Zhejiang University, Hangzhou 310027, China
2
ZJU-Hangzhou Global Scientific and Technological Innovation Center, Zhejiang University, Hangzhou 311200, China
3
Shanghai Institute of Satellite Engineering, Shanghai 201109, China
4
Jiaxing Key Laboratory of Photonic Sensing & Intelligent Imaging, Jiaxing 314000, China
5
Intelligent Optics & Photonics Research Center, Jiaxing Research Institute Zhejiang University, Jiaxing 314000, China
*
Author to whom correspondence should be addressed.
Atmosphere 2022, 13(9), 1384; https://doi.org/10.3390/atmos13091384
Submission received: 28 July 2022 / Revised: 15 August 2022 / Accepted: 26 August 2022 / Published: 29 August 2022
(This article belongs to the Section Aerosols)

Abstract

:
China is developing the High-precision Greenhouse gases Monitoring Satellite (HGMS), carrying a high-spectral-resolution lidar (HSRL) for aerosol vertical profiles and imaging grating spectrometers for CO2 measurements at the same time. By providing simultaneous evaluation of the aerosol scattering effect, HGMS would reduce the bias of the XCO2 retrievals from the passive sensor. In this work, we propose a method to reduce aerosol-induced bias in XCO2 retrievals for the future HGMS mission based on the correlation analysis among simulated radiance, XCO2 bias, and aerosol optical depth (AOD) ratio. We exercise the method with the Orbiting Carbon Observatory-2 (OCO-2) XCO2 retrievals and AOD ratio inferred from the OCO-2 O2 A-band aerosol parameters at 755 nm and the Cloud–Aerosol Lidar with Orthogonal Polarization (CALIOP) AOD at 532 nm at several Total Carbon Column Observing Network (TCCON) sites in Europe. The results showed that 80% of measurements from OCO-2 were improved, and data from six TCCON sites show an average of 2.6 ppm reduction in mean bias and a 68% improvement in accuracy. We demonstrate the advantage of fused active–passive observation of the HGMS for more accurate global XCO2 measurements in the future.

1. Introduction

Based on the requirement of the National Civil Space Infrastructure and Long-term Development Plan (2015–2025), China will launch the High-precision Greenhouse gases Monitoring Satellite (HGMS) in 2023. HGMS was designed as an active–passive fusion satellite to carry a wide-swath greenhouse gases monitor (GGM) and an aerosol–cloud high-spectral-resolution lidar (ACHSRL) on a single platform in a 705 km solar synchronous orbit [1,2]. GGM covers a swath of more than 100 km and has four spectral bands: 0.753–0.768 μm, 1.595–1.625 μm, 2.04–2.08 μm, and 2.275–2.325 μm, which are used for the aerosol loads estimation and the column-averaged CO2 and CH4 mole fraction (XCO2 and XCH4) detection, respectively. ACHSRL consists of a 532 nm iodine-based HSRL and a 1064 nm Mie lidar. By separating the Mie backscatter return from the Rayleigh backscatter return, ACHSRL could directly provide the extinction profile of aerosols and clouds [3]. Furthermore, the optical properties at 532 nm and 1064 nm would help to better identify the aerosol types [4]. These two payloads make HGMS capable of simultaneous observation for CO2 and aerosol information [5]. Since the influence of aerosols on XCO2 retrieval is a recognized major issue [6,7,8,9,10], the simultaneous observation of XCO2 and aerosols of HGMS will help estimate and correct the aerosol scattering effect on XCO2 retrievals, which will significantly improve the accuracy of global XCO2 measurements.
The mechanism and function of GGM’s first three bands are similar to those of the Orbiting Carbon Observatory-2 (OCO-2). The added fourth band at 2.276–2.325 µm is specially used for CH4 retrievals, the second important greenhouse gas. Currently, the mainstream CO2 retrieval method for passive satellite remote sensing is the full physical retrieval algorithm [9,11,12,13]. The aerosol information used in the algorithm is mainly from models or reanalysis data. However, aerosols have complex subtypes and relatively rapid temporal and spatial variability in the real atmosphere [14,15]. Aerosol observations from passive remote sensing can only retrieve the column aerosol optical depth (AOD) with restricted accuracy [16,17]. As a result, their ability to correct the aerosol scattering effect on XCO2 retrievals is very limited. For example, it is easy to misestimate the influence on the upper atmospheric structure due to the lack of the ability to distinguish the upper layer (stratospheric) thin clouds and aerosols essentially. The data shows about a 20% deviation between cloud screen products from OCO-2 and cloud mask products from the Moderate Resolution Imaging Spectrometer (MODIS) [6]. The clouds and aerosols located at higher altitudes, such as stratospheric sulfate from volcanic eruptions, could cause larger XCO2 bias (100 ppm) even with a small optical depth (such as AOD < 0.3) [7,18].
Studies of XCO2 retrieval accuracy improvement considering the aerosol-induced bias have been performed. Before the launch of OCO-2, Wunch et al. proposed a method to reduce the XCO2 systematic bias for the Greenhouse gases Observing SATellite (GOSAT) by constructing the correction equation between the XCO2 and several parameters (albedo, pressure, and airmass) that may cause XCO2 bias [10]. Furthermore, Guerlet et al. found the correlations between the XCO2 bias and aerosol parameters from GOSAT and tried to reduce the interference from the aerosol scattering effect on account of the linear regression fit [7]. Based on Wunch and Guerlet’s research, O’Dell et al. constructed multilinear bias-correction fitting to the surface pressure, CO2 gradient, and AOD to obtain the correlation with XCO2, which helped to improve the OCO-2 XCO2 retrieval and update the OCO-2 version 8 algorithm [18]. All the corrections aforementioned used the Total Carbon Column Observing Network (TCCON) measurements as the truth [19,20,21]. These XCO2 bias-corrected algorithms almost rely on the predetermined aerosol models, which are not completely consistent with the real atmosphere.
Active remote sensing, such as lidar, can acquire the real-time and accurate vertical structure of the atmosphere [3,22,23]. Using lidar-observed aerosol products to improve XCO2 accuracy has been studied and proved to be effective. Uchino et al. reduced GOSAT retrieval errors by introducing the vertical distribution of aerosols and cirrus observed by ground-based lidar to the retrieval algorithm [24]. Taylor et al. compared OCO-2 observation with collocated Cloud–Aerosol Lidar with Orthogonal Polarization (CALIOP) measurements to analyze the performance of the OCO-2 cloud screening algorithm and found the advantage of CALIOP in identifying high, thin clouds [25]. Merrelli et al. used the CALIOP-derived aerosol products as prior for the OCO-2 physical retrieval [26], which suggested that the OCO-2/CALIOP tight formation flying in A-train is excellent for studying active–passive lidar/OCO-2 retrievals.
HGMS will carry active–passive fusion instruments, providing the real-time aerosol data to correct aerosol-induced bias in XCO2 retrievals. To demonstrate the capability of HGMS to improve the XCO2 accuracy, we used the OCO-2 and CALIOP data due to their tight flying formation in the following analysis [26].
In this study, a retrieved XCO2 accuracy improvement method for HGMS was proposed based on the correlation analysis between XCO2 bias and AOD ratio. Solar radiances at the weak CO2 absorption band under different atmospheric scenes were firstly simulated for HGMS to quantify the aerosol scattering effect to the radiance bias. Then, the OCO-2 data and CALIOP AOD were matched to exercise our proposed method. Two-year (2015–2016) OCO-2/CALIOP measurements over land surface and six TCCON sites in Europe were analyzed. With reference to Guerlet et al. and O’Dell et al. [7,18], segmented linear correlation was calculated between XCO2 bias and AOD ratio at four data sites, which accounts for the impact from different aerosol species, and used to correct XCO2 at independent validation sites with close background at Bremen and Paris. Due to the complexity of aerosol layer height and subtype distribution, this paper mainly focuses on the seasonal and regional influence of CO2 measurements caused by aerosol loadings in the troposphere.
After a brief introduction of HGMS, the construction method of HGMS’s solar radiance simulation and the substituted datasets of HGMS used in this work are described in Section 2. Section 3 first presents the simulated radiance distribution and analyzes the correlation between the simulated radiance bias and AOD ratio. Then, the seasonal and regional correlation between XCO2 bias and AOD ratio are presented. The correlation analysis results are used to improve the OCO-2 official bias-corrected products for six TCCON sites in Europe. Section 4 provides the conclusion of this study.

2. Materials and Methods

2.1. Basic Configuration of HGMS

The detailed parameters of ACHSRL and GGM of HGMS can be found in Table 1. GGM is essentially an imaging grating spectrometer, collecting high-resolution spectra of reflected sunlight in the O2 A-band, the weak and strong CO2 absorption band, and the CH4 absorption band. Similar to OCO-2, the O2 A-band spectrum has a high sensitivity to changes in the light path [27]. It will be primarily used to determine the surface pressure and screen for optically thick clouds [25]. Furthermore, it can be used to quantify the AOD and vertical distribution of aerosols for soundings that are sufficiently optically thin (AOD < 0.3) to yield accurate estimates of XCO2. The other active lidar instrument, ACHSRL, consists of a 532 nm HSRL and a 1064 Mie scattering lidar, which is capable of aerosol and cloud optical properties profiling. The real-time aerosol and cloud information provided by ACHSRL could replace the model data during the GGM XCO2 retrieval and help to accurately estimate the aerosol scattering effect.
The configurations and detection mechanism of GGM and ACHSRL are similar to OCO-2 and CALIOP. All these instruments are loaded on their corresponding satellite in a 705 km solar synchronous orbit (CALIOP lowered its orbit from 705 km to 688 km after September 2018) [18,28]. Their orbit inclination and repeat cycles are all 98.2° and 16 days. ACHSRL and CALIOP are both dual-wavelength lidar with Nd:YAG lasers that emit pulses at 532 and 1064 nm. Their pulse energy at 532 nm is 110 mJ and the repeat frequencies are both close to 20 Hz. Furthermore, their divergence angles are 0.1 mrad and the field of view of ACHSRL is 0.2 mrad, while the field of view of CALIOP is 0.13 mrad [28]. Considering the same orbit altitude, inclination angle, and laser repeat frequency, the global laser footprint coverage of ACHSRL and CALIOP could be assumed to be the same. Besides the instrument configuration, ACHSRL and CALIOP are both polarized lidar collecting the backscattering light to retrieve the optical properties of aerosols and clouds. The spectral ranges of OCO-2 are 0.76–0.77, 1.59–1.62, and 2.04–2.08 μm, respectively, which are nearly the same as GGM’s first three bands. The spatial resolution of OCO-2 is 1.29 × 2.25 km2, while GGM’S resolution is close to 3 × 3 km2 [18]. Both of their retrieval methods are the full physical retrieval algorithm. Though CALIOP and OCO-2 are loaded on different satellites, they could still acquire the collocated near-simultaneous measurements due to the A-train formation. Since HGMS has not been launched yet, simulated GGM and ACHSRL observations are substituted for OCO-2 and CALIOP datasets.
The product information can be found in Table 2. The XCO2 from OCO-2 Standard products were used to analyze the correlation between the bias of XCO2 and aerosol loads, while the Lite products were used to validate the feasibility of the correction algorithm because of its better agreement with the ground base stations and model data, which is not in the Standard products [29]. The TCCON XCO2 products are used as the truth [30,31]. The aerosol information analyzed was the aerosol parameters from the O2 A-band at 755 nm provided by OCO-2 Standard products and the observed extinction coefficient profiles at 532 nm provided by CALIOP; the CALIOP AOD is assumed as the true value. The cloud–aerosol discrimination (CAD) scores of CALIOP products were used to filter data with cloud contamination and low confidence [28].

2.2. Radiance Simulation for HGMS

To better understand the HGMS XCO2 bias caused by aerosol scattering effect, simulation for the solar radiance at the HGMS’s weak CO2 absorption band under different aerosol optical properties was calculated. The radiative transfer model used for atmosphere absorption is the Line-By-Line Radiative Transfer Model (LBLRTM v12.9) [32,33]; multiple scattering simulation is calculated by library for radiative transfer (libRadtran v2.0.4) with the discrete ordinates radiative transfer solvers (DISORT) [34,35,36]; solar irradiance spectrum is simulated by the Kurucz compilation [37]. Other parameters chosen in this study mainly refer to Mao’s work, which can be found in Table 3 [8]. The vertical resolution of the atmosphere profiles (temperature, pressure, etc.) is 100 m and the surface reflectance is 0.2, considering the real situation of the European land surface [38].
Besides the molecular absorption and the surface reflection, the aerosol scattering effect is also considered. Different aerosol scenes, including marine, rural, and urban aerosols, were tested, which were the main aerosol types over European land [39,40,41]. All the aerosols are distributed below 2 km. All aerosol models are taken from LOWTRAN and the aerosol optical properties are calculated by Mie theory and the Henyey–Greenstein phase function according to Mao et al. [8], which can be found in Table 4. All the models and parameters are input in the libRadtran software with Fortran program [35].

2.3. XCO2 and AOD Datasets for HGMS

2.3.1. Data Screening and Fusion

To acquire matched comparison data pairs among OCO-2, CALIOP, and TCCON and eliminate the errors caused by temporal and spatial mismatch, a few steps were taken to screen the observations from these instruments. The dates when there are no data available for satellite footprints within a 200 km radius from the TCCON site, and the minimum distance between OCO-2 and CALIOP footprint is over 10 km, are removed. The number of available dates for different sites is counted in Table A1. CALIOP absolute CAD score higher than 70 indicates high confidence from its official data introduction (positive values signify clouds, whereas negative values signify aerosols) [4]. Since the cloud existence would lead to XCO2 retrieval failure, CALIOP profiles with all the CAD scores lower than −70 are retained for further analysis. After orbit registration, cloud screening, and quality control, the remaining OCO-2 soundings and CALIOP footprints can be placed into pairs for the correlation comparison. The procedure is to choose the CALIOP footprints as the center point and draw a rectangle around the point along the orbit direction as the fusion range. The rectangle size is 5 km (the horizontal resolution of CALIOP L2 products) along the orbit direction and 10 km across the orbit direction, which is displayed in the dotted red circle in Figure 1 [42]. The soundings of OCO-2 located in the fusion range are averaged. The averaged AOD of OCO-2 and CALIOP are denoted as τ755 and τ532, respectively. The averaged XCO2 of OCO-2 is denoted as XCO2, OCO2. These fused data are used in Section 3.2 for data correlation analysis.
Six TCCON sites in Europe, including Bialystok, Bremen, Garmisch, Karlsruhe, Orleans, and Paris, are analyzed in this work [10,21]. Bremen and Paris are used to validate the XCO2 accuracy improvement method, while the other four sites are used to analyze the correlation. Due to the high cloud fraction, large surface snow cover, and limited direct sunlight in winter (December–January–February, DJF) and spring (March–April–May, MAM), most of the TCCON and OCO-2 measurements are in summer (June–July–August, JJA) and fall (September–October–November, SON). Liang et al. compared the nadir mode OCO-2 retrievals and TCCON measurements by restricting the period within an hour and the area within ±2° latitude and ±2.5° longitude of the TCCON sites. The comparison found that the effect of averaging kernels changes could be neglected with the differences less than ±0.021 ppm on average [43]. In this work, a 200 km radius area from the TCCON sites and ±1 h period measurements were also chosen to obtain sufficient data and guarantee spacetime correlation (as shown in Figure 1). Synergistic data in 2015 and 2016 were analyzed because CALIOP has experienced lower energy laser shots within and near the South Atlantic Anomaly region since 2017 and it also lowered its orbit from 705 km to 688 km to resume formation with CloudSat in 2018. Furthermore, the TCCON data availability in Europe is higher in 2015 and 2016 [44]. Totally, 2625 comparison pairs were obtained, covering 111 days over two years.

2.3.2. Data Segmentation

For different TCCON sites, different seasons, and different true values of AOD are considered. The data are divided into four seasons (DJF, MAM, JJA, SON), and the data in the same season are divided into four regions based on the τ532. When τ532 = 0, it means the atmosphere should be clean without any aerosol. When τ532 ≠ 0, it means there are aerosols in the atmosphere, which may lead to bias of XCO2 measurements. Aben et al. found typical errors in the SCanning Imaging Absorption spectroMeter for Atmospheric CHartographY (SCIAMACHY) CO2 column for nadir observations over land surfaces with AOD range from 0.1 to 0.3 [45]. Crisp et al. identified 0.3 as a reasonable AOD threshold below which to attempt XCO2 retrievals [46]. Therefore, we divided the AOD into several regions based on the distribution of the values, which are 0 < τ532 ≤ 0.1, 0.1 < τ532 ≤ 0.3, τ532 > 0.3. Here, we use XXX (aod1, aod2) to describe a case in the correlation at a TCCON site. For instance, MAM (0.1, 0.3] means the correlation is analyzed when 0.1 < τ532 ≤ 0.3 and in spring.
The Ångström exponent law was introduced here to estimate the AOD ratio between different wavelengths:
τ/τ0 = (λ/λ0)k.
λ is the wavelength and τ is the corresponding AOD to be estimated; λ0 is the reference wavelength and τ0 is the corresponding reference AOD; k is the well-known Ångström exponent [47]. For example, Amiridis et al. used AErosol RObotic NETwork (AERONET) and CALIOP measurements to establish a multi-wavelength aerosol optical databases at 355, 532, 1064, 1570, and 2050 nm by employing the Ångström exponent law [47]. Holben et al. also used AERONET measurements to calculate the Ångström exponent at 440/870 nm [48]. Based on their calculation and statistics, the approximation of k755/532 with the same size distribution and aerosol types could be roughly estimated as 1.6, which means the τ755/532 could be estimated as 0.57. If τ755/532 is far beyond or lower than 0.57, the τ755 may have a large bias compared with its true value, assuming the τ532 is accurate. The XCO2 bias ∆XCO2 can be expressed as [19]
XCO2 = (XCO2, TCCONXCO2, OCO2)/XCO2, TCCON × 100%,
where XCO2, TCCON is the TCCON XCO2 measurements averaged within ±1 h when OCO-2 passed through the site (assumed the amount of OCO-2 measurements is N and TCCON measurements is M). Meanwhile, the uncertainty of XCO2 bias δXCO2 is calculated as [49]
XCO2/∆XCO2)2 = (δXCO2, TCCON/XCO2, TCCON)2 + (δXCO2, OCO2/XCO2, OCO2)2.
δXCO2, TCCON and δXCO2, OCO2 are the uncertainty of TCCON and OCO-2 XCO2 measurements, computed as [19]
( δ X C O 2 ,   O C O 2 ) 2 = 1 N 1 i = 1 N ( X C O 2 ,   O C O 2 i X ¯ C O 2 ,   O C O 2 ) ,
( δ X C O 2 ,   T C C O N ) 2 = 1 M 1 i = 1 M ( X C O 2 ,   T C C O N i X ¯ C O 2 ,   T C C O N ) .
Further analysis between τ755/532 and ∆XCO2 is described in Section 3.2.

3. Results and Discussion

3.1. Radiance Simulation Analysis

The radiance distributions of top of atmosphere in HGMS’s weak CO2 absorption band under different aerosol types and optical depths are illustrated in Figure 2. In Figure 2a, the radiance changes under different aerosol scenes with the same AOD. The radiation deviations caused by aerosols can be positive and negative compared with the clear sky, and the largest and smallest deviations occur under urban and rural aerosol scenes. Figure 2b shows the radiance distribution influenced by different AOD under urban aerosol scene. It is easy to see that the radiation deviations increased when the AOD increased.
Here, we defined R as the ratio between the online and offline radiance (ron and roff), R = ron/roff. The offline was set to 6382.7 cm−1 and the online was from 6300 cm−1 to 6340 cm−1. The ratio change means the ratio variation when the XCO2 increased by 1% under PBL, which is △R = (R’ − R)/R × 100%. R’ means the ratio when XCO2 is 363.6 ppm. The ratio change distribution under different aerosol scenes can be found in Figure 3a–d. The AODs set in Figure 3a–c are all 0.5. The distribution indicates that the aerosol existence leads to positive ratio change, which means that the aerosol scattering effect could lengthen the optical path of CO2 detection, and it resulted in an overestimated XCO2 retrieval. Additionally, the largest △R was introduced under the marine aerosol scene while the smallest one occurred under urban aerosol scene with the same AOD of 0.5. Figure 3e–h shows the radiance change when the XCO2 increased by 1% with different AOD under urban aerosol scene. In the clear sky (Figure 3h), the peak value (−0.4%) of the spectrum is the largest compared with the aerosol scenes, which means that the detection sensitivity would be the largest and the relative error would be the smallest where there is no aerosol load. The sensitivity is only half of that in the clear sky when the AOD is 0.5 in Figure 3g, which would demand higher signal-to-noise ratio (SNR). The bias in Figure 3e–h means the peak value bias of the spectrum compared with the clear sky.
The relationship between radiance bias and AOD ratio (τ755/532) was also simulated to analyze the XCO2 bias influenced by the inaccurate OCO-2 aerosol model products. The results can be found in Figure 4. As described in Section 2, the AOD at 532 nm was assumed to be the true value. The radiance bias ∆roff in Figure 4a was calculated at the offline wavelength with different AOD input, ∆roff = roff,755roff,532. The intercept on the x-axis is 0.57, which is assumed to be the accurate value of τ755/532. When τ755/532 = 0.57, it could be recognized that there is no AOD bias at 755 nm and the radiance bias is correspondingly zero. Figure 4a shows the radiance bias influenced by the τ755/532 with different AOD values (0.05–0.4) at 532 nm. The higher τ755/532 could lead to higher radiance bias. Figure 4b shows the relationship between median radiance changes ∆rm and τ755/532. ∆rm are the medians of the spectrum radiance difference when the AOD inputs are set to 755 nm and 532 nm, respectively.
rm = median|(rband,755rband,532)/rband,532 × 100%|.
rband means the radiance spectrum distributed from 6300 to 6400 cm−1. The symbol of median || means calculating the median of radiance changes. The ∆rm in Figure 4b are all positive, which means the XCO2 retrieval would be overestimated. The larger the τ755/532, the more the results are overestimated. Additionally, when the AOD at 532 nm is lower than 0.1, the radiance bias and changes have a linear correlation with the τ755/532. With the increase of AOD, the correlation gradually turns into a nonlinear form, which is caused by complicated multiple scattering effect with higher aerosol loading. Consequently, in the high-AOD scenes, the quality of HGMS’s CO2 retrieval is worrying. With limited dynamic detection range and spectral resolution, invalid retrieval would occur for HGMS if there is large bias for the AOD estimation.

3.2. Correlation Analysis

The simulation results in Figure 4 indicate that the correlation between AOD ratio and HGMS’s radiance bias could be linear or near-linear. As for the correlation between HGMS’s XCO2 bias and AOD ratio, the correlation coefficients with the division of different cases are presented in Table 5, where LD means lack of data and LC means low correlation in terms of a linear relationship (correlation coefficient R < 0.5). NDA means the amount of data available after the screening method described in Section 2. NHC means the amount of data of high correlation (R > 0.5).
The results indicate that about half of the cases (22/48, regardless of the LD cases) have a certain correlation between ∆XCO2 and τ755/532 (τ755). For the 22 cases with a certain correlation, we assume the linear regression equation is expressed as y × 100 = b + a × x, where x means τ755/532 (τ755) and y means ∆XCO2. The results are displayed in Table 6.
Correlation plots of the six cases are displayed in Figure 5 and Figure 6, with τ532 = 0 and τ532 ≠ 0, respectively.

3.3. Validation

The difference in surface albedo may lead to systematic errors in CO2 retrievals [18]. To make the correlation analysis as independent of surface albedo as possible, the weekly averaged surface albedo at the six TCCON sites from April to December (no data available for validation in January, February, or March at Bremen and Paris) were calculated based on the MODIS MCD43C3 products at band 6 (1.628–1.652 μm) [50]. The results can be seen in Figure A1. The weekly averaged albedo at Bialystok is 0.42 ± 0.01, while the albedo at Bremen is 0.44 ± 0.01. The weekly averaged albedo at Orleans is 0.25 ± 0.02, and the albedo is 0.23 ± 0.01 at Paris. We decide to use the correlation analysis results at Bialystok to improve the Bremen observation, and we used the Orleans results to improve the Paris observation due to their similar weekly albedo change.
According to Equation (2) and the linear regression between ∆XCO2 and τ755/532, it can be derived that
XCO2 = b + a × τ755/532.
The optimization OCO-2 observations are denoted as XCO2, NEW, which is used to replace the XCO2, TCCON in Equation (2) because it is assumed that the TCCON measurements are the true values, and then we can obtain the optimization XCO2, NEW after approximate linear regression correction by calculating
(XCO2, NEWXCO2, OCO2)/XCO2, NEW × 100% = b + a × τ755/532.
The XCO2, NEW can be expressed as
XCO2, NEW = XCO2, OCO2/(1 − ba×τ755/532) × 100%.
The optimization XCO2, NEW is obtained based on the linear regression coefficient a and b selected from Table 6. The OCO-2 Lite products are used to validate the feasibility of the correction algorithm. Figure 7 shows the comparison results between OCO-2 and TCCON measurements for Bremen and Paris sites.
The OCO-2 measurements include official Standard, Lite products, and new results after linear regression correction, noted as XCO2, Std, XCO2, Lite, and XCO2, NEW. The ∆XCO2 in Figure 7 equals XCO2, i minus XCO2, TCCON (i = Std, Lite, and New), while the legends in Figure 7 represent the statistical results of the absolute value of the XCO2 difference. The ∆XCO2 in Figure 7 are divided into several regions, and the heights of different colored histograms represent the counts distributed in the corresponding region.
In Figure 7a–c, the difference between XCO2, OCO2 and XCO2, TCCON is decreased from 1.68 ± 0.91 ppm to 0.99 ± 1.03 ppm on September 7 2015, with 0 < τ532 ≤ 0.1, while the difference between XCO2, Lite and XCO2, TCCON is 0.77 ± 0.7 ppm. In addition, the difference between XCO2, OCO2 and XCO2, TCCON is decreased from 2.59 ± 1.36 ppm to 0.95 ± 1.12 ppm on November 28 2016, with 0.1 < τ532 ≤ 0.3, while the difference between XCO2, Lite and is 1.12 ± 1.16 ppm. In Figure 7g–i, the difference between XCO2, OCO2 and XCO2, TCCON is decreased from 7.98 ± 3.23 ppm to 4.99 ± 3.63 ppm on May 21 2015, with 0.1 < τ532 ≤ 0.3, while the difference between XCO2, Lite and XCO2, TCCON is 4.78 ± 2.75 ppm. The difference between XCO2, OCO2 and XCO2, TCCON is decreased from 3.81 ± 0.81 ppm to 1.64 ± 1.58 ppm on September 10 2015, with τ532 = 0, while the difference between XCO2, Lite and XCO2, TCCON is 2.26 ± 0.77 ppm.
A total of 16 days (7 days at Bremen and 9 days at Paris) of linear regression correction results are presented in Figure 8. Here, the mean bias (MB) of the XCO2 is introduced to further indicate the overall bias between the OCO-2 and TCCON measurements. MB can be expressed as
MB = 1 N N | X C O 2 ,   TCCON X C O 2 ,   O C O 2 | .
Both XCO2, OCO2 and XCO2, TCCON are daily averaged results. The correlation coefficients R are improved from 0.96 to 0.98 for Bremen and from 0.65 to 0.68 for Paris, while the MB is decreased from 2.39 ppm to 0.89 ppm for Bremen and from 5.86 ppm to 1.07 ppm for Paris. Table 7 gives more detailed information about the comparisons among TCCON, OCO-2, and optimization results of XCO2 for Bremen and Paris during 2015–2016.
Linear regression correction results of all six TCCON sites are analyzed and presented in Figure 9. The correlation coefficient R is improved from 0.76 to 0.81, and the MB is decreased from 3.81 ppm to 1.21 ppm.

3.4. Discussion

In Table 5, the overall data utilization is 49.4%, which means that about half of the data is correlated based on the division. The statistical histogram is presented in Figure 10. The rectangular bars with different line types in the same season mean different TCCON sites. The single bar is the data amount for one site in a specific season. The amount of the green pattern is equal to the NHC in Table 5, while the orange pattern represents data with a lower correlation (NDA-NHC). For different seasons, it is clear that the total amounts (NDA) in DJF and MAM are lower than the other seasons, which contribute only 24.3% of the NDA due to the high cloud fraction and lack of reflected solar radiation from the sun in winter and spring in the Northern Hemisphere. As for the cloud fraction over land surface in Europe in 2016, Chen et al. found that the highest clear-sky fraction occurs in summer (55.1%), while the lowest occurs in winter (29.5%). The spring and fall have close values, which are 39.3% and 40.0%, respectively [6]. In winter, the Bialystok site has the lowest NDA because of its lower solar radiance located at high latitude. Additionally, NDA would be different since the TCCON data availability was different, considering the instrument failure and weather conditions. Based on the official TCCON datasets, there are 394 days of available data for Bialystok, 278 days for Garmisch, 267 days for Karlsruhe, and 415 days for Orleans in 2015–2016, respectively [51].
As for the cases with lower correlation in Table 6, the possible reasons are the following: a) there is limited NDA in this case, such as the MAM(0, 0.1] in Bialystok, the MAM(0.1, 0.3] in Garmisch, and the MAM(0.3, ∞) in Orleans, the NDA of which are, respectively, 12, 10, and 12. Lower NDA may lead to lower correlation because a single datum can significantly influence the correlation; b) though the NDA is enough, the data are from a single date, which made the results reflect just one day’s situation. For instance, the data in MAM(0.1, 0.3] in Karlsruhe came from only 7 May 2015 and 2 May 2016. The NDA is 44 on May 2 while the NDA is 4 on May 7, which represents the case analysis for the single day’s correlation and cannot reflect the seasonal characterization; c) the same aerosol load in the same season can lead to different errors to △XCO2 if the aerosol subtype and vertical distribution are quite different, which is consistent with the finding in Section 3.1.
Figure 5 indicates that most of the τ755 is less than 0.4, which is a reasonable range of the AOD from the OCO-2 aerosol model. The bigger the τ755 is, the bigger the △XCO2 is for most data, revealing that the increase of τ755 can lead to the increase of △XCO2 when τ532 = 0 in summer at these two sites. In Figure 6, most of the τ755/532 is larger than 0.57, which shows that the τ755 is overestimated, assuming the τ532 is the true value. The correlation is consistent with the simulation results in Figure 4, which means that the linear regression fitting equation applied here is reasonable. Similar to the trends in Figure 5, it can also indicate that the increase of τ755/532 can lead to the increase of △XCO2 when τ532 ≠ 0 in specific seasons and at specific sites in Figure 6.
As for the validation section, the distribution of stacked histograms in Figure 7 shows significant trends where most Standard products are smaller than the assumed true value of TCCON measurements. In comparison, the Lite products and the optimization results are closer. Though there are some cases where the optimization results are not as improved as the Lite products, the deviations are usually less than 0.2 ppm, which is far less than the uncertainty of each product. It is worth mentioning that the uncertainty of new optimization results is occasionally larger than the Standard products. This is mainly related to the linear regression equation. Not all the residuals are small enough in the fitting regions, which means that if the τ755/532 near the Bremen site is close to zero or greater than 3, the optimization results may not be satisfied. The 16 days’ validation results in Figure 8 and Table 7 show that 79.7% (424/532) data pairs are successfully improved using this method. For the Bremen site, the differences between the average XCO2 from TCCON sites and OCO-2 measurements passing nearby regions is decreased from 2.39 ± 2.48 ppm to 0.89 ± 2.41 ppm, with a 1.5 ppm mean bias decrease. For the Paris site, the result is decreased from 5.86 ± 3.70 ppm to 1.07 ± 3.16 ppm, with a 4.79 ppm mean bias decrease. The accuracy is improved by 51.9% and 81.7%, respectively. Though the SDCO2 of optimization data is not decreased in JJA and SON at the Bremen site from Table 7, it is still acceptable because the increments are less than 0.2 ppm (0.19 and 0.06 ppm, respectively). As for the total six TCCON site measurements in Figure 9, the differences between the averaged XCO2 from TCCON sites are decreased from 3.81 ± 2.62 ppm to 1.21 ± 1.89 ppm, with a 2.6 ppm mean bias decrease and 68.2% accuracy improvement. In general, the statistical results verify the feasibility of the proposed method to reduce aerosol-induced bias in XCO2 retrievals for the future HGMS.

4. Conclusions

In this study, an XCO2 accuracy improvement method was proposed based on the radiative transfer simulation and two-year measurements from OCO-2, CALIOP, and six TCCON sites to demonstrate the advantage of active–passive fusion observation. The method was designed for HGMS XCO2 observations in the future.
Simulations for the solar radiance at the HGMS’s weak CO2 absorption band under different aerosol parameters (AOD, subtype) were calculated. For AOD, the radiance change was −0.4% when XCO2 increased by 1% in the clear sky, but the change would be half of that when the AOD is 0.5; for aerosol subtypes, the largest radiance ratio change was introduced under the marine aerosol scene while the smallest one occurred under the urban aerosol. The lower radiance change represents lower detection sensitivity, which means that higher SNR was needed for the detector and a larger XCO2 bias would occur in the urban scenes with large AOD. In addition, a near-linear correlation was found between the radiance bias and AOD ratio (τ755/532), which provided the main idea of our XCO2 accuracy improvement method.
Under the guidance of radiance simulation, the correlations between XCO2 bias and AOD ratio of HGMS were analyzed. OCO-2 and CALIOP measurements were substituted for the corresponding datasets of HGMS. First, a screening and fusion method was used to acquire proper data for correlation analysis. Under specific TCCON site, season, and AOD, the XCO2 bias and AOD ratio have a linear correlation. Among 2625 comparison pairs, 49.4% of them show a certain linear correlation (R > 0.56). For different seasons, the minimum NDA and NHC occur in winter due to the high cloud fraction and lack of solar radiation in the Northern Hemisphere. A lookup table of the linear regression equation coefficient was summarized for 22 different cases at four TCCON sites. An XCO2 accuracy improvement method was proposed based on the lookup table and linear regression equations. The method was then applied to the other two TCCON sites (Bremen and Paris). Under the comparison of OCO-2 official Lite products, 80.0% (424/532) of the data used to validate the linear regression correction method are successfully improved. Similar to the four experimental TCCON sites, there is no retrieved XCO2 that is bias-corrected in winter because of the cloud existence and solar radiance issues for the two validation sites. For the rest of data in spring, summer, and fall, the XCO2 bias is decreased from 2.39 ± 2.48 ppm to 0.89 ± 2.41 ppm for Bremen and from 5.86 ± 3.70 ppm to 1.07 ± 3.16 ppm for Paris. All six TCCON data show a 2.6 ppm mean bias decrease and 68.2% accuracy improvement. The correlation coefficient was also improved from 0.76 to 0.81. The feasibility of the XCO2 accuracy improvement method by reducing aerosol-induced bias was demonstrated.
For future work, more measured data with different aerosol layer heights and subtypes could be considered, as aerosol profiling and classification are the advantage of lidar. Since the XCO2 data used in this study are from 2015–2016 over land surface in Europe, more data in different years and near different TCCON sites around the world could be analyzed for further verification, which will be useful to update the optimization method applied to HGMS in the future and will help to obtain more accurate XCO2 observations globally.

Author Contributions

Conceptualization, J.K. and D.L.; methodology, J.K. and D.L.; data curation, J.K., Y.S. and S.W.; writing—original draft preparation, J.K.; writing—review and editing, C.D., S.C. and D.L.; funding acquisition, D.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Excellent Young Scientist Program of the Zhejiang Provincial Natural Science Foundation of China, grant number LR19D050001; the Fundamental Research Funds for the Zhejiang Provincial Universities, grant number 2021XZZX019; the State Key Laboratory of Modern Optical Instrumentation Innovation Program, grant number MOI2021ZD01; a project supported by Scientific Research Fund of Zhejiang University, grant number XY2021050.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The OCO-2 L2 Standard products (V10r) are available at https://disc.gsfc.nasa.gov/datasets/OCO2_L2_Standard_10/summary?keywords=OCO2, accessed on 1 May 2021. The OCO-2 L2 Lite products (V10r) are available at https://disc.gsfc.nasa.gov/datasets/OCO2_L2_Lite_FP_10r/summary?keywords=OCO2, accessed on 1 October, 2021. The TCCON products (GGG2014) are available at https://tccondata.org/2014, accessed on 1 May 2021. The CALIOP L2 aerosol/cloud profile products (V4-20) are available at https://subset.larc.nasa.gov/calipso/login.php, accessed on 11 December, 2021. The Line-By-Line molecular absorption databases are available at https://hitran.org/, accessed on 1 October 2021. The MODIS MCD43C3 products are available at https://ladsweb.modaps.eosdis.nasa.gov/search/order/1/MCD43C3--6, accessed on 1 October 2021.

Acknowledgments

The authors would like to thank the science teams of OCO-2, CALIOP, TCCON, and MODIS for providing excellent and accessible data products used in this investigation.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Band 6 of MODIS is similar to the weak CO2 absorption band of OCO-2, and albedo data with relatively high SNR can be obtained from this band. The MCD43C3 products are available at https://ladsweb.modaps.eosdis.nasa.gov/search/order/1/MCD43C3--6, accessed on 1 October 2021 [50].
Figure A1. Weekly averaged surface albedo at the six TCCON sites based on the MODIS MCD43C3 products at band 6 (1.628–1.652 μm).
Figure A1. Weekly averaged surface albedo at the six TCCON sites based on the MODIS MCD43C3 products at band 6 (1.628–1.652 μm).
Atmosphere 13 01384 g0a1

Appendix B

Table A1. Summary of the available dates of different sites used for correlation analysis.
Table A1. Summary of the available dates of different sites used for correlation analysis.
Site (Location)Date
Bialystok
(53.23° N, 23.025° E)
26 Feb 2015
23 Mar 2015
25 Mar 2015
30 Mar 2015
12 May 2015
11 Jun 2015
13 Jun 2015
6 Jul 2015
20 Jul 2015
31 Jul 2015
1 Sep 2015
22 Sep 2015
10 Oct 2015
25 Nov 2015
18 Mar 2016
25 Mar 2016
19 Apr 2016
26 Apr 2016
3 May 2016
21 May 2016
28 May 2016
4 Jun 2016
22 Jun 2016
29 Jun 2016
24 Jul 2016
31 Jul 2016
7 Aug 2016
1 Sep 2016
8 Sep 2016
19 Sep 2016
22 Nov 2016
29 Nov 2016
Karlsruhe
(49.10° N, 8.44° E)
7 May 2015
17 Jun 2015
13 Aug 2015
21 Sep 2015
24 Nov 2015
19 Dec 2015
26 Dec 2015
20 Jan 2016
2 May 2016
10 Jun 2016
5 Jul 2016
7 Jul 2016
21 Jul 2016
30 Jul 2016
1 Aug 2016
6 Aug 2016
8 Aug 2016
31 Aug 2016
7 Sep 2016
14 Sep 2016
16 Oct 2016
Bremen
(53.10° N, 8.85° E)
13 Aug 2015
7 Sep 2015
11 Apr 2016
2 May 2016
24 Aug 2016
31 Aug 2016
28 Nov 2016
Garmisch
(47.48° N, 11.06° E)
18 Feb 2015
18 May 2015
10 Jun 2015
17 Jun 2015
5 Jul 2015
6 Aug 2015
13 Aug 2015
21 Sep 2015
10 Nov 2015
24 Nov 2015
20 May 2016
28 Jun 2016
5 Jul 2016
31 Aug 2016
7 Sep 2016
27 Oct 2016
3 Nov 2016
Orleans
(47.97° N, 2.113° E)
21 Feb 2015
23 Mar 2015
25 Mar 2015
1 Apr 2015
10 May 2015
19 May 2015
21 May 2015
13 Jun 2015
15 Jul 2015
22 Jul 2015
29 Jul 2015
10 Sep 2015
24 Sep 2015
12 Oct 2015
19 Oct 2015
26 Oct 2015
4 Dec 2015
22 Dec 2015
20 Mar 2016
27 Mar 2016
10 Apr 2016
28 Apr 2016
5 May 2016
6 Jun 2016
24 Jun 2016
8 Jul 2016
15 Jul 2016
9 Aug 2016
27 Aug 2016
3 Sep 2016
10 Sep 2016
28 Sep 2016
19 Oct 2016
30 Oct 2016
6 Nov 2016
Paris
(48.846° N, 2.356° E)
12 May 2015
21 May 2015
15 Jul 2015
10 Sep 2015
24 Sep 2015
26 Oct 2015
24 Jun 2016
10 Sep 2016
28 Sep 2016

References

  1. Liu, Y.; Wang, J.; Che, K.; Cai, Z.; Yang, D.; Wu, l. Satellite remote sensing of greenhouse gases: Progress and trends. Natl. Remote Sens. Bull. 2021, 25, 53–64. [Google Scholar] [CrossRef]
  2. Han, G.; Ma, X.; Liang, A.; Zhang, T.; Zhao, Y.; Zhang, M.; Gong, W. Performance Evaluation for China’s Planned CO2-IPDA. Remote Sens. 2017, 9, 768. [Google Scholar] [CrossRef]
  3. Wang, N.; Zhang, K.; Shen, X.; Wang, Y.; Li, J.; Li, C.; Mao, J.; Malinka, A.; Zhao, C.; Russell, L.M.; et al. Dual-field-of-view high-spectral-resolution lidar: Simultaneous profiling of aerosol and water cloud to study aerosol-cloud interaction. Proc. Natl. Acad. Sci. USA 2022, 119, e2110756119. [Google Scholar] [CrossRef] [PubMed]
  4. Hostetler, C.A.; Lee, K.-P.; Ferrare, R.A.; Trepte, C.R.; Hu, Y.; Vaughan, M.A.; Winker, D.M.; Omar, A.H.; Kittaka, C.; Rogers, R.R.; et al. The CALIPSO Automated Aerosol Classification and Lidar Ratio Selection Algorithm. J. Atmos. Ocean. Technol. 2009, 26, 1994–2014. [Google Scholar] [CrossRef]
  5. Friedlingstein, P.; Cox, P.; Betts, R.; Bopp, L.; Von Bloh, W.; Brovkin, V.; Cadule, P.; Doney, S.; Eby, M.; Fung, I.; et al. Climate-carbon cycle feedback analysis: Results from the (CMIP)-M-4 model intercomparison. J. Clim. 2006, 19, 3337–3353. [Google Scholar] [CrossRef]
  6. Chen, S.; Wang, S.; Su, L.; Dong, C.; Ke, J.; Zheng, Z.; Cheng, C.; Tong, B.; Liu, D. Optimization of the OCO-2 Cloud Screening Algorithm and Evaluation against MODIS and TCCON Measurements over Land Surfaces in Europe and Japan. Adv. Atmos. Sci. 2020, 37, 387–398. [Google Scholar] [CrossRef]
  7. Guerlet, S.; Butz, A.; Schepers, D.; Basu, S.; Hasekamp, O.P.; Kuze, A.; Yokota, T.; Blavier, J.F.; Deutscher, N.M.; Griffith, D.W.T.; et al. Impact of aerosol and thin cirrus on retrieving and validating XCO2 from GOSAT shortwave infrared measurements. J. Geophys. Res. Atmos. 2013, 118, 4887–4905. [Google Scholar] [CrossRef]
  8. Mao, J.; Kawa, S.R. Sensitivity studies for space-based measurement of atmospheric total column carbon dioxide by reflected sunlight. Appl. Opt. 2004, 43, 914–927. [Google Scholar] [CrossRef]
  9. O’Dell, C.W.; Connor, B.; Boesch, H.; O’Brien, D.; Frankenberg, C.; Castano, R.; Christi, M.; Crisp, D.; Eldering, A.; Fisher, B.; et al. The ACOS CO2 retrieval algorithm—Part 1: Description and validation against synthetic observations. Atmos. Meas. Tech. 2012, 5, 99–121. [Google Scholar] [CrossRef]
  10. Wunch, D.; Wennberg, P.O.; Toon, G.C.; Connor, B.J.; Fisher, B.; Osterman, G.B.; Frankenberg, C.; Mandrake, L.; O’Dell, C.; Ahonen, P.; et al. A method for evaluating bias in global measurements of CO2 total columns from space. Atmos. Chem. Phys. 2011, 11, 12317–12337. [Google Scholar] [CrossRef] [Green Version]
  11. 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]
  12. Yoshida, Y.; Kikuchi, N.; Morino, I.; Uchino, O.; Oshchepkov, S.; Bril, A.; Saeki, T.; Schutgens, N.; Toon, G.C.; Wunch, D.; et al. 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]
  13. Crisp, D.; Fisher, B.M.; O’Dell, C.; Frankenberg, C.; Basilio, R.; Boesch, H.; Brown, L.R.; Castano, R.; Connor, B.; Deutscher, N.M.; et al. The ACOS CO2 retrieval algorithm—Part II: Global X-CO2 data characterization. Atmos. Meas. Tech. 2012, 5, 687–707. [Google Scholar] [CrossRef]
  14. Wei, J.; Peng, Y.; Mahmood, R.; Sun, L.; Guo, J. Intercomparison in spatial distributions and temporal trends derived from multi-source satellite aerosol products. Atmos. Chem. Phys. 2019, 19, 7183–7207. [Google Scholar] [CrossRef]
  15. Hu, Z.; Jin, Q.; Ma, Y.; Pu, B.; Dong, W. Temporal evolution of aerosols and their extreme events in polluted Asian regions during Terra’s 20-year observations. Remote Sens. Environ. 2021, 263, 112541. [Google Scholar] [CrossRef]
  16. Stier, P. Limitations of passive remote sensing to constrain global cloud condensation nuclei. Atmos. Chem. Phys. 2016, 16, 6595–6607. [Google Scholar] [CrossRef]
  17. Christensen, M.W.; Neubauer, D.; Poulsen, C.; Thomas, G.; Grainger, R.G. Unveiling aerosol–cloud interactions—Part 1: Cloud contamination in satellite products enhances the aerosol indirect forcing estimate. Atmos. Chem. Phys. 2017, 17, 13151–13164. [Google Scholar] [CrossRef]
  18. 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]
  19. Zhou, M.; Dils, B.; Wang, P.; Detmers, R.; Yoshida, Y.; O’Dell, C.W.; Feist, D.G.; Velazco, V.A.; Schneider, M.; De Mazière, M. 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]
  20. Yang, D.; Boesch, H.; Liu, Y.; Somkuti, P.; Cai, Z.; Chen, X.; Di Noia, A.; Lin, C.; Lu, N.; Lyu, D.; et al. Toward High Precision XCO2 Retrievals From TanSat Observations: Retrieval Improvement and Validation Against TCCON Measurements. J. Geophys. Res. Atmos 2020, 125, e2020JD032794. [Google Scholar] [CrossRef]
  21. Wunch, D.; Wennberg, P.O.; Osterman, G.; Fisher, B.; Naylor, B.; Roehl, C.M.; O’Dell, C.; Mandrake, L.; Viatte, C.; Kiel, M.; et al. Comparisons of the Orbiting Carbon Observatory-2 (OCO-2) XCO2 measurements with TCCON. Atmos. Meas. Tech. 2017, 10, 2209–2238. [Google Scholar] [CrossRef]
  22. Liu, D.; Zheng, Z.; Chen, W.; Wang, Z.; Li, W.; Ke, J.; Zhang, Y.; Chen, S.; Cheng, C.; Wang, S. Performance estimation of space-borne high-spectral-resolution lidar for cloud and aerosol optical properties at 532 nm. Opt. Express 2019, 27, A481–A494. [Google Scholar] [CrossRef] [PubMed]
  23. Ke, J.; Sun, Y.; Dong, C.; Zhang, X.; Wang, Z.; Lyu, L.; Zhu, W.; Ansmann, A.; Su, L.; Bu, L.; et al. Development of China’s first space-borne aerosol-cloud high-spectral-resolution lidar: Retrieval algorithm and airborne demonstration. PhotoniX 2022, 3, 17. [Google Scholar] [CrossRef]
  24. Uchino, O.; Kikuchi, N.; Sakai, T.; Morino, I.; Yoshida, Y.; Nagai, T.; Shimizu, A.; Shibata, T.; Yamazaki, A.; Uchiyama, A.; et al. Influence of aerosols and thin cirrus clouds on the GOSAT-observed CO2: A case study over Tsukuba. Atmos. Chem. Phys. 2012, 12, 3393–3404. [Google Scholar] [CrossRef]
  25. Taylor, T.E.; O’Dell, C.W.; Frankenberg, C.; Partain, P.T.; Cronk, H.Q.; Savtchenko, A.; Nelson, R.R.; Rosenthal, E.J.; Chang, A.Y.; Fisher, B.; et al. 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]
  26. Merrelli, A.J.; Bennartz, R.; O’Dell, C. Evaluating XCO2 Retrievals from OCO-2 with CALIOP-Derived Aerosol Priors. In Agu Fall Meeting Abstracts, Proceedings of the the 2017 AGU Fall Meeting, New Orleans, LA, USA, 11–15 December 2017; American Geophysical Union: Washington, DC, USA, 2017. [Google Scholar]
  27. Oyafuso, F.; Payne, V.H.; Drouin, B.J.; Devi, V.M.; Benner, D.C.; Sung, K.; Yu, S.; Gordon, I.E.; Kochanov, R.; Tan, Y.; et al. High accuracy absorption coefficients for the Orbiting Carbon Observatory-2 (OCO-2) mission: Validation of updated carbon dioxide cross-sections using atmospheric spectra. J. Quant. Spectrosc. Radiat. Transf. 2017, 203, 213–223. [Google Scholar] [CrossRef]
  28. Powell, K.A.; Hu, Y.; Omar, A.; Vaughan, M.A.; Winker, D.M.; Liu, Z.; Hunt, W.H.; Young, S.A. Overview of the CALIPSO Mission and CALIOP Data Processing Algorithms. J. Atmos. Ocean. Technol. 2009, 26, 2310–2323. [Google Scholar] [CrossRef]
  29. OCO-2 Science Team; Gunson, M.; Eldering, A. OCO-2 Level 2 Geolocated XCO2 Retrievals Results, Physical Model, Retrospective Processing V10r, version 10r; Goddard Earth Sciences Data and Information Services Center: Greenbelt, MD, USA, 2020. [CrossRef]
  30. Deng, A.; Yu, T.; Cheng, T.; Gu, X.; Zheng, F.; Guo, H. Intercomparison of Carbon Dioxide Products Retrieved from GOSAT Short-Wavelength Infrared Spectra for Three Years (2010–2012). Atmosphere 2016, 7, 109. [Google Scholar] [CrossRef]
  31. Tadić, J.; Biraud, S. An Approach to Estimate Atmospheric Greenhouse Gas Total Columns Mole Fraction from Partial Column Sampling. Atmosphere 2018, 9, 247. [Google Scholar] [CrossRef] [Green Version]
  32. Clough, S.A.; Shephard, M.W.; Mlawer, E.J.; Delamere, J.S.; Iacono, M.J.; Cady-Pereira, K.; Boukabara, S.; Brown, P.D. Atmospheric radiative transfer modeling: A summary of the AER codes. J. Quant. Spectrosc. Radiat. Transf. 2005, 91, 233–244. [Google Scholar] [CrossRef]
  33. Fomin, B.; Falaleeva, V. A Polarized Atmospheric Radiative Transfer Model for Calculations of Spectra of the Stokes Parameters of Shortwave Radiation Based on the Line-by-Line and Monte Carlo Methods. Atmosphere 2012, 3, 451–467. [Google Scholar] [CrossRef]
  34. Stamnes, K.; Tsay, S.C.; Wiscombe, W.; Laszlo, I. DISORT, a General-Purpose Fortran Program for Discrete-Ordinate-Method Radiative Transfer in Scattering and Emitting Layered Media: Documentation of Methodology, version 1.1; NTRS: Chicago, IL, USA, 2000. [Google Scholar]
  35. Emde, C.; Buras-Schnell, R.; Kylling, A.; Mayer, B.; Gasteiger, J.; Hamann, U.; Kylling, J.; Richter, B.; Pause, C.; Dowling, T. The libRadtran software package for radiative transfer calculations (version 2.0.1). Geosci. Model Dev. 2016, 9, 1647–1672. [Google Scholar] [CrossRef]
  36. Stamnes, K.; Tsay, S.C.; Wiscombe, W.J.; Jayaweera, K. Numerically stable algorithm for discrete-ordinate-method radiative transfer in multiple scattering and emitting layered media. Appl. Opt. 1988, 27, 2502–2509. [Google Scholar] [CrossRef] [PubMed]
  37. Kurucz, R.L. Synthetic Infrared Spectra. In Symposium-International Astronomical Union; Cambridge University Press: Singapore, 1994; Volume 154, p. 523. [Google Scholar]
  38. Wang, M.; Fan, X.; Li, X.; Liu, Q.; Qu, Y. Estimation of Land Surface Albedo from MODIS and VIIRS Data: A Multi-Sensor Strategy Based on the Direct Estimation Algorithm and Statistical-Based Temporal Filter. Remote Sens. 2020, 12, 4131. [Google Scholar] [CrossRef]
  39. Logothetis, S.A.; Salamalikis, V.; Kazantzidis, A. Aerosol classification in Europe, Middle East, North Africa and Arabian Peninsula based on AERONET Version 3—ScienceDirect. Atmos. Res. 2020, 239, 104893. [Google Scholar] [CrossRef]
  40. Garofalide, S.; Postolachi, C.; Cocean, A.; Cocean, G.; Motrescu, I.; Cocean, I.; Munteanu, B.S.; Prelipceanu, M.; Gurlui, S.; Leontie, L. Saharan Dust Storm Aerosol Characterization of the Event (9 to 13 May 2020) over European AERONET Sites. Atmosphere 2022, 13, 493. [Google Scholar] [CrossRef]
  41. Milinevsky, G.; Miatselskaya, N.; Grytsai, A.; Danylevsky, V.; Bril, A.; Chaikovsky, A.; Yukhymchuk, Y.; Wang, Y.; Liptuga, A.; Kyslyi, V.; et al. Atmospheric Aerosol Distribution in 2016–2017 over the Eastern European Region Based on the GEOS-Chem Model. Atmosphere 2020, 11, 722. [Google Scholar] [CrossRef]
  42. Qiu, R.; Han, G.; Ma, X.; Xu, H.; Shi, T.; Zhang, M. A Comparison of OCO-2 SIF, MODIS GPP, and GOSIF Data from Gross Primary Production (GPP) Estimation and Seasonal Cycles in North America. Remote Sens. 2020, 12, 258. [Google Scholar] [CrossRef]
  43. Liang, A.; Gong, W.; Han, G.; Xiang, C. Comparison of Satellite-Observed XCO2 from GOSAT, OCO-2, and Ground-Based TCCON. Remote Sens. 2017, 9, 1033. [Google Scholar] [CrossRef] [Green Version]
  44. Chen, X.; Liu, Y.; Yang, D.; Cai, Z.; Chen, H.; Wang, M. A Theoretical Analysis for Improving Aerosol-Induced CO2 Retrieval Uncertainties Over Land Based on TanSat Nadir Observations Under Clear Sky Conditions. Remote Sens. 2019, 11, 1061. [Google Scholar] [CrossRef]
  45. 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]
  46. Crisp, D.; Atlas, R.M.; Breon, F.M.; Brown, L.R.; Burrows, J.P.; Ciais, P.; Connor, B.J.; Doney, S.C.; Fung, I.Y.; Jacob, D.J.; et al. The Orbiting Carbon Observatory (OCO) mission. Adv. Space Res. 2004, 34, 700–709. [Google Scholar] [CrossRef]
  47. Amiridis, V.; Marinou, E.; Tsekeri, A.; Wandinger, U.; Schwarz, A.; Giannakaki, E.; Mamouri, R.; Kokkalis, P.; Binietoglou, I.; Solomos, S.; et al. LIVAS: A 3-D multi-wavelength aerosol/cloud database based on CALIPSO and EARLINET. Atmos. Chem. Phys. 2015, 15, 7127–7153. [Google Scholar] [CrossRef]
  48. Holben, B.N.; Tanré, D.; Smirnov, A.; Eck, T.F.; Slutsker, I.; Abuhassan, N.; Newcomb, W.W.; Schafer, J.S.; Chatenet, B.; Lavenu, F.; et al. An emerging ground-based aerosol climatology: Aerosol optical depth from AERONET. J. Geophys. Res. Atmos. 2001, 106, 12067–12097. [Google Scholar] [CrossRef]
  49. Russell, P.B.; Swissler, T.J.; McCormick, M.P. Methodology for error analysis and simulation of lidar aerosol measurements. Appl. Opt. 1979, 18, 3783–3797. [Google Scholar] [CrossRef] [PubMed]
  50. Schaaf, C.; Wang, Z. MCD43C3 MODIS/Terra+Aqua BRDF/Albedo Albedo Daily L3 Global 0.05Deg CMG V006, version 6; NASA EOSDIS Land Processes DAAC: Sioux Falls, SD, USA, 2015. [CrossRef]
  51. Wunch, D.; Toon, G.C.; Blavier, J.; Washenfelder, R.A.; Notholt, J.; Connor, B.J.; Griffith, D.; Sherlock, V.; Wennberg, P.O. The Total Carbon Column Observing Network. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2011, 369, 2087–2112. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Distribution of selected TCCON (white pentagrams) sites in Europe. The orange lines show OCO-2 footprints passing over nearby TCCON sites, and the green lines show the CALIOP footprints. The dotted red circle shows the data fusion range between OCO-2 and CALIOP footprints.
Figure 1. Distribution of selected TCCON (white pentagrams) sites in Europe. The orange lines show OCO-2 footprints passing over nearby TCCON sites, and the green lines show the CALIOP footprints. The dotted red circle shows the data fusion range between OCO-2 and CALIOP footprints.
Atmosphere 13 01384 g001
Figure 2. Radiance distribution influenced by aerosols of different types and AOD at HGMS’s weak CO2 absorption band. (a) CO2 absorption spectrum in the clear sky, urban, marine, and rural aerosols; (b) CO2 absorption spectrum with different AOD under urban aerosol scene.
Figure 2. Radiance distribution influenced by aerosols of different types and AOD at HGMS’s weak CO2 absorption band. (a) CO2 absorption spectrum in the clear sky, urban, marine, and rural aerosols; (b) CO2 absorption spectrum with different AOD under urban aerosol scene.
Atmosphere 13 01384 g002
Figure 3. Radiance ratio (on/off) changes when XCO2 increased by 1% under different atmospheric scenes: (ad) maritime, rural, urban, clear sky; radiance changes when XCO2 increased by 1% with different AOD (eh) AOD = 0.1, 0.3, 0.5, 0.7, and clear sky.
Figure 3. Radiance ratio (on/off) changes when XCO2 increased by 1% under different atmospheric scenes: (ad) maritime, rural, urban, clear sky; radiance changes when XCO2 increased by 1% with different AOD (eh) AOD = 0.1, 0.3, 0.5, 0.7, and clear sky.
Atmosphere 13 01384 g003
Figure 4. Radiance bias and median radiance changes with different τ755/532. (a) Dots in different colors mean different AOD at 532 nm. The shaded block indicates the variation range of ∆roff, and the dashed line indicates the polynomial fitting curves between ∆roff and τ755/532. (b) △rm distribution with different τ755/532.
Figure 4. Radiance bias and median radiance changes with different τ755/532. (a) Dots in different colors mean different AOD at 532 nm. The shaded block indicates the variation range of ∆roff, and the dashed line indicates the polynomial fitting curves between ∆roff and τ755/532. (b) △rm distribution with different τ755/532.
Atmosphere 13 01384 g004
Figure 5. Correlation analysis between ∆XCO2 and τ755 when τ532 = 0: (a) JJA season at Bialystok site; (b) JJA season at Garmisch site. The blue shaded areas represent the standard error (1σ and 2σ).
Figure 5. Correlation analysis between ∆XCO2 and τ755 when τ532 = 0: (a) JJA season at Bialystok site; (b) JJA season at Garmisch site. The blue shaded areas represent the standard error (1σ and 2σ).
Atmosphere 13 01384 g005
Figure 6. Correlation analysis between ∆XCO2 and τ755/532 when τ532 ≠ 0: (a) SON(0, 0.1] at Bialystok site; (b) MAM(0, 0.1] at Garmisch site; (c) MAM(0.1, 0.3] at Bialystok site; (d) JJA(0.1, 0.3] at Garmisch site. The blue shaded areas represent the standard error (1σ and 2σ).
Figure 6. Correlation analysis between ∆XCO2 and τ755/532 when τ532 ≠ 0: (a) SON(0, 0.1] at Bialystok site; (b) MAM(0, 0.1] at Garmisch site; (c) MAM(0.1, 0.3] at Bialystok site; (d) JJA(0.1, 0.3] at Garmisch site. The blue shaded areas represent the standard error (1σ and 2σ).
Atmosphere 13 01384 g006
Figure 7. Stacked histogram of XCO2 measurement difference between the XCO2, i and XCO2, TCCON (i = Std, New, and Lite). The orange, blue, and pink histograms represent the Standard, new optimization, and Lite products. Comparison results: (ac) 7 September 2015 at Bremen; (df) 28 November 2016 at Bremen; (gi) 21 May 2015 at Paris; (jl) 10 September 2015 at Paris.
Figure 7. Stacked histogram of XCO2 measurement difference between the XCO2, i and XCO2, TCCON (i = Std, New, and Lite). The orange, blue, and pink histograms represent the Standard, new optimization, and Lite products. Comparison results: (ac) 7 September 2015 at Bremen; (df) 28 November 2016 at Bremen; (gi) 21 May 2015 at Paris; (jl) 10 September 2015 at Paris.
Atmosphere 13 01384 g007
Figure 8. Scatterplots of seasonal daily average XCO2 from OCO-2 versus collocated TCCON observations. The orange and blue dots are respective results before and after linear regression correction. The dashed line is the one-to-one ratio line. The upward, downward, left, and right triangle points represent the spring, summer, fall, and winter results. (a) 7 days of results at Bremen; (b) 9 days of results at Paris.
Figure 8. Scatterplots of seasonal daily average XCO2 from OCO-2 versus collocated TCCON observations. The orange and blue dots are respective results before and after linear regression correction. The dashed line is the one-to-one ratio line. The upward, downward, left, and right triangle points represent the spring, summer, fall, and winter results. (a) 7 days of results at Bremen; (b) 9 days of results at Paris.
Atmosphere 13 01384 g008
Figure 9. Scatterplots of seasonal daily average XCO2 from OCO-2 versus collocated TCCON observations. The dashed line is the one-to-one ratio line. The red, green, blue, and yellow points represent the spring, summer, fall, and winter results. (a) Results before linear regression correction; (b) results after linear regression correction.
Figure 9. Scatterplots of seasonal daily average XCO2 from OCO-2 versus collocated TCCON observations. The dashed line is the one-to-one ratio line. The red, green, blue, and yellow points represent the spring, summer, fall, and winter results. (a) Results before linear regression correction; (b) results after linear regression correction.
Atmosphere 13 01384 g009
Figure 10. Statistical histogram of data amounts in different seasons and at different TCCON sites. The ordinate is the amount of data. The green pattern means the data has a certain correlation (NHC), while the orange pattern has a minor correlation. The orthogonal straight line indicates the data in Bialystok. The wave line in the vertical direction shows the data in Garmisch, the obliquely crossed straight line shows the data in Karlsruhe, and the oblique straight line represents the data in Orleans.
Figure 10. Statistical histogram of data amounts in different seasons and at different TCCON sites. The ordinate is the amount of data. The green pattern means the data has a certain correlation (NHC), while the orange pattern has a minor correlation. The orthogonal straight line indicates the data in Bialystok. The wave line in the vertical direction shows the data in Garmisch, the obliquely crossed straight line shows the data in Karlsruhe, and the oblique straight line represents the data in Orleans.
Atmosphere 13 01384 g010
Table 1. Parameters of ACHSRL and GGM of HGMS.
Table 1. Parameters of ACHSRL and GGM of HGMS.
ACHSRLGGM
ParameterValueParameterValue
Laser wavelength532/1064 nmSpectral range0.753–0.768/1.595–1.625/
2.04–2.08/2.275–2.325 μm
Pulse energy150/110 mJ
Frequency20 HzSpectral resolution<0.04/0.07/0.09/0.1 nm
Laser bandwidth<100 MHzSignal-to-noise ratio>350/340/230/200
Filter bandwidth<2 GHzSpatial resolution<3 km
Divergence angle0.1 mradSwath>100 km
Field of view0.2 mrad
Table 2. Datasets used for the correlation analysis between XCO2 bias and aerosol scattering effect.
Table 2. Datasets used for the correlation analysis between XCO2 bias and aerosol scattering effect.
InstrumentProduct NameParameters
XCO2OCO-2L2 Standard V10rXCO2
L2 Lite FP V10rBias corrected XCO2
TCCONGGG2014XCO2
AerosolOCO-2L2 Standard V10rAerosol parameters at 755 nm
CALIOPL2 5km A/CPro V4-20Extinction coefficient at 532 nm
CAD score
Table 3. Input parameters of radiance simulation for HGMS.
Table 3. Input parameters of radiance simulation for HGMS.
ParametersValue
CO2 level360 ppm
Atmosphere profiles1976 U.S. Standard Atmosphere
Lambertian surface reflectance0.2
Solar zenith angle32°
Satellite viewing anglenadir
Solar irradiance spectrumKurucz compilation
Spectral band range6300–6400 cm−1
Table 4. Optical properties of different aerosols used in the radiance simulation.
Table 4. Optical properties of different aerosols used in the radiance simulation.
AerosolSingle Scattering AlbedoAsymmetryVisibility/km
Marine0.980.7223
Rural0.810.645
Urban0.520.635
Table 5. The correlation coefficients between HGMS’s XCO2 bias and AOD ratio at different seasons, different τ532, and different TCCON Sites. LD means lack of data. LC means low correlation.
Table 5. The correlation coefficients between HGMS’s XCO2 bias and AOD ratio at different seasons, different τ532, and different TCCON Sites. LD means lack of data. LC means low correlation.
SiteSeasonτ532 = 0(0, 0.1](0.1, 0.3](0.3, ∞)NHC/NDAData Utilization
BialystokDJFLDLD0.62LD17/1894.4%
MAMLCLC0.560.5895/19947.7%
JJA0.560.62LC0.64221/28577.5%
SONLC0.61LCLC122/27544.4%
GarmischDJF0.63LDLDLD50/50100%
MAMLD0.57LCLD29/3974.5%
JJA0.58LC0.560.69192/26771.9%
SONLCLCLCLC0/2810
KarlsruheDJF0.58LCLDLD42/9444.7%
MAMLC0.64LCLD45/9348.3%
JJALC0.60LCLC70/18737.4%
SONLCLC0.57LD106/25341.9%
OrleansDJFLDLD0.79LD10/1758.8%
MAMLCLD0.64LC35/12927.1%
JJA0.62LCLC0.70120/19661.2%
SON0.59LC0.64LD143/24238.5%
TOTAL 1297/262549.4%
Table 6. Lookup table of linear regression equation coefficients of 22 cases.
Table 6. Lookup table of linear regression equation coefficients of 22 cases.
SiteCasebaSiteCaseba
BialystokJJAτ532 = 00.094.71GarmischDJFτ532 = 00.3711.8
JJA(0, 0.1]0.490.53JJAτ532 = 00.185.44
SON(0, 0.1]0.640.47MAM(0, 0.1]0.0570.39
DJF(0.1, 0.3]0.760.27JJA(0.1, 0.3]0.140.34
MAM(0.1, 0.3]−0.200.59JJA(0.3, ∞)0.0621.58
MAM(0.3, ∞)0.620.90OrleansJJAτ532 = 00.185.44
JJA(0.3, ∞)0.281.89SONτ532 = 00.427.91
KarlsruheDJFτ532 = 00.159.98DJF(0.1, 0.3]0.411.27
MAM(0, 0.1]0.150.48MAM(0.1, 0.3]0.270.35
JJA(0, 0.1]0.180.26SON(0.1, 0.3]0.130.52
SON(0.1, 0.3]0.480.51JJA(0.3, ∞)0.301.18
Table 7. Summary of the XCO2 among TCCON, OCO-2, and optimization results for Bremen and Paris during 2015–2016. SDCO2 is the standard deviation of XCO2.
Table 7. Summary of the XCO2 among TCCON, OCO-2, and optimization results for Bremen and Paris during 2015–2016. SDCO2 is the standard deviation of XCO2.
SiteSeasonTCCONOCO-2 DataOptimization OCO-2 Data
XCO2CountXCO2SDCO2RCountXCO2SDCO2R
BremenMAM403.8217400.461.670.9615401.991.220.98
JJA398.53155397.121.90106398.352.09
SON399.30152396.903.86140398.743.92
ParisMAM403.0758396.304.100.6523399.963.170.68
JJA399.5153394.694.6443400.524.28
SON398.6797392.692.3697397.562.02
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ke, J.; Wang, S.; Chen, S.; Dong, C.; Sun, Y.; Liu, D. Retrieved XCO2 Accuracy Improvement by Reducing Aerosol-Induced Bias for China’s Future High-Precision Greenhouse Gases Monitoring Satellite Mission. Atmosphere 2022, 13, 1384. https://doi.org/10.3390/atmos13091384

AMA Style

Ke J, Wang S, Chen S, Dong C, Sun Y, Liu D. Retrieved XCO2 Accuracy Improvement by Reducing Aerosol-Induced Bias for China’s Future High-Precision Greenhouse Gases Monitoring Satellite Mission. Atmosphere. 2022; 13(9):1384. https://doi.org/10.3390/atmos13091384

Chicago/Turabian Style

Ke, Ju, Shuaibo Wang, Sijie Chen, Changzhe Dong, Yingshan Sun, and Dong Liu. 2022. "Retrieved XCO2 Accuracy Improvement by Reducing Aerosol-Induced Bias for China’s Future High-Precision Greenhouse Gases Monitoring Satellite Mission" Atmosphere 13, no. 9: 1384. https://doi.org/10.3390/atmos13091384

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