Advanced Search
Article Contents

The First Global Map of Atmospheric Ammonia (NH3) as Observed by the HIRAS/FY-3D Satellite


doi: 10.1007/s00376-023-3059-9

  • Atmospheric ammonia (NH3) is a chemically active trace gas that plays an important role in the atmospheric environment and climate change. Satellite remote sensing is a powerful technique to monitor NH3 concentration based on the absorption lines of NH3 in the thermal infrared region. In this study, we establish a retrieval algorithm to derive the NH3 column from the Hyperspectral Infrared Atmospheric Sounder (HIRAS) onboard the Chinese FengYun (FY)-3D satellite and present the first atmospheric NH3 column global map observed by the HIRAS instrument. The HIRAS observations can well capture NH3 hotspots around the world, e.g., India, West Africa, and East China, where large NH3 emissions exist. The HIRAS NH3 columns are also compared to the space-based Infrared Atmospheric Sounding Interferometer (IASI) measurements, and we find that the two instruments observe a consistent NH3 global distribution, with correlation coefficient (R) values of 0.28–0.73. Finally, some remaining issues about the HIRAS NH3 retrieval are discussed.
    摘要: 氨气(NH3)在地球大气中是一种化学性质活泼的微量气体,在大气环境和气候变化中发挥着重要作用。卫星遥感是借助于NH3在热红外谱段的特征吸收线,反演获得NH3浓度的有效技术。在本研究中,基于最优估计理论建立了一套NH3柱浓度的反演算法,成功应用于中国风云-3D卫星上的高光谱红外大气探测仪(HIRAS),获得了HIRAS仪器观测到的大气NH3柱全球分布图。风云-3D卫星观测可以很好地捕捉全球NH3高值区,例如印度、西非、中国东部等存在大量NH3排放的地区。风云-3D卫星与欧洲红外大气探测干涉仪卫星(IASI)测量的NH3结果具有良好的一致性,相关系数 (R) 值为0.28–0.73。最后,本文针对风云-3D/HIRAS卫星NH3反演目前仍存在一些问题进行讨论,并提出了未来风云卫星发展方向。
  • 加载中
  • Figure 1.  (a) A typical HIRAS observed spectra in the long wavelength range (648–1136 cm–1; spectral resolution of 0.625 cm–1) at 18.72°N, 70.50°E, for a satellite zenith angle of 3.27° and a solar zenith angle of 45.25°. The NH3 retrieval window (960–970 cm–1) is marked by a red shadow. (b) The observed (Obs) and simulated (Sim) spectra, and (c) the residual of the fitted spectrum (Obs – Sim).

    Figure 2.  (a) The transmittances of the main species (CO2, H2O, O3, and NH3) in the NH3 retrieval window, and (b) a zoom window in the transmittance range between 0.965 and 1.00 to establish a better view of NH3 absorption lines.

    Figure 3.  The CAMS model global simulated NH3 mole fraction near the surface in January, April, July, and October between 2015 and 2020.

    Figure 4.  Typical NH3 vertical profiles from the CAMS model over land and ocean. (a) Typical averaging kernel matrix (solid lines; lines are colored with different altitudes; in units of ppb (ppb)−1 and column averaging kernel scaled with 0.1 (red dashed line; in units of (molecules cm–2) (molecules cm–2)−1 of the HIRAS NH3 retrieval over (b) land and (c) ocean.

    Figure 5.  The retrieved NH3 columns on 30 January 2020 from the HIRAS observations under clear-sky conditions during daytime (a) and nighttime (c), together with the differences between the retrieved columns and a priori columns (re-ap) during daytime (b) and nighttime (d), respectively.

    Figure 6.  The NH3 column monthly means in January 2020 observed by (a) HIRAS/FY3D and (b) IASI/MetopB re-gridded onto 1° × 1° (latitude × longitude). The scatter plots between HIRAS and IASI NH3 columns in (c) North America, (d) Europe, (e) East China, (f) South America, (g) Africa, and (h) India. In each scatter plot, the dots are colored according to the data density. The red dashed line is the linear regression. The solid black line is the one-to-one line. R is the Pearson correlation coefficient and N is the number of the data points.

    Figure 7.  Same as Fig. 6, but for July 2020.

    Figure 8.  (a) The NH3 columns, (b), the RMSE, and (c) the DOF for all convergent retrievals derived from the HIRAS spectra in January 2020.

    Figure 9.  The RMSE (left) and DOF (right) as a function of the latitude bin (every 10 degrees) as derived from all the HIRAS NH3 retrievals (grey) and the retrievals after quality filtering (orange) in January 2020. The error bar is the standard deviation.

    Figure 10.  Typical NH3 retrievals over the ocean using the (a) HIRAS observed spectrum and (b) the IASI observed spectrum. Note that the a priori NH3 in the title denotes the NH3 mole fraction near the surface.

    Table 1.  The state vector ($ \mathbf{x} $) together with their prior and variance (1σ) in the HIRAS NH3 retrieval algorithm.

    Retrieved parameters ($ \mathbf{x} $) Prior 1σ
    NH3 column CAMS model SD of CAMS monthly means
    O3 column CAMS model 50%
    CO2 column Carbon Tracker 10%
    H2O column ERA5 reanalysis data 100%
    Surface temperature ERA5 reanalysis data 3%
    Spectral shift 0 10%
    DownLoad: CSV

    Table 2.  The relative mean and std of the difference between HIRAS and IASI measurements [(HIRAS−IASI)/IASI × 100%] over six regions with high NH3 columns in January and July 2020, together with their Pearson correlation coefficients (R), p-values, and mean DOFs.

    Region (latitude/longitude)
    North America Europe East China South America Africa India
    [15°N, 55°N]/
    [120°W, 80°W]
    [30°N, 60°N]/
    [20°W, 30°E]
    [15°N, 45°N]/
    [100°E, 120°E]
    [60°S, 0°]/
    [80°W, 30°W]
    [0°, 18°N]/
    [20°W, 60°E]
    [10°N, 35°N]/
    [60°E, 90°E]
    Jan-2020 Mean (%) –2.0 –3.1 –17.2 35.6 –5.5 –5.2
    Std (%) 52.9 65.1 65.1 82.4 51.7 46.6
    R 0.28 0.31 0.69 0.55 0.68 0.73
    P-value <0.001 <0.001 <0.001 <0.001 <0.001 <0.001
    DOF 0.47 0.48 0.61 0.53 0.49 0.57
    Jul-2020 Mean (%) 42.5 11.4 –3.0 –9.9 –11.0 –3.6
    Std (%) 78.7 53.4 82.3 71.5 51.8 60.8
    R 0.50 0.32 0.31 0.32 0.47 0.58
    P-value <0.001 <0.001 <0.001 <0.001 <0.001 <0.001
    DOF 0.57 0.66 0.44 0.45 0.42 0.54
    DownLoad: CSV
  • NOAA., 1976: U.S. Standard Atmosphere, 1976. United States Committee on Extension to the Standard Atmosphere, 227 pp.
    Bao, Z. E., and Coauthors, 2021: Effects of NH3 on secondary aerosol formation from toluene/NOx photo-oxidation in different O3 formation regimes. Atmos. Environ., 261, 118603, https://doi.org/10.1016/j.atmosenv.2021.118603.
    Behera, S. N., M. Sharma, V. P. Aneja, and R. Balasubramanian, 2013: Ammonia in the atmosphere: A review on emission sources, atmospheric chemistry and deposition on terrestrial bodies. Environmental Science and Pollution Research, 20, 8092−8131, https://doi.org/10.1007/s11356-013-2051-9.
    Bouwman, A. F., D. S. Lee, W. A. H. Asman, F. J. Dentener, K. W. Van Der Hoek, and J. G. J. Olivier, 1997: A global high-resolution emission inventory for ammonia. Global Biogeochemical Cycles, 11, 561−587, https://doi.org/10.1029/97GB02266.
    Clarisse, L., C. Clerbaux, F. Dentener, D. Hurtmans, and P.-F. Coheur, 2009: Global ammonia distribution derived from infrared satellite observations. Nature Geoscience, 2, 479−483, https://doi.org/10.1038/ngeo551.
    Clarisse, L., and Coauthors, 2010: Satellite monitoring of ammonia: A case study of the San Joaquin Valley. J. Geophys. Res. Atmos., 115, D13302, https://doi.org/10.1029/2009JD 013291.
    Crippa, M., and Coauthors, 2020: High resolution temporal profiles in the emissions database for global atmospheric research. Scientific Data, 7, 121, https://doi.org/10.1038/s41597-020-0462-2.
    Dammers, E., and Coauthors, 2016: An evaluation of IASI-NH3 with ground-based Fourier transform infrared spectroscopy measurements. Atmospheric Chemistry and Physics, 16 , 10 351−10 368, https://doi.org/10.5194/acp-16-10351-2016.
    Dammers, E., and Coauthors, 2017: Validation of the CrIS fast physical NH3 retrieval with ground-based FTIR. Atmospheric Measurement Techniques, 10, 2645−2667, https://doi.org/10.5194/amt-10-2645-2017.
    De Wachter, E., N. Kumps, A. C. Vandaele, B. Langerock, and M. De Mazière, 2017: Retrieval and validation of MetOp/IASI methane. Atmospheric Measurement Techniques, 10, 4623−4638, https://doi.org/10.5194/amt-10-4623-2017.
    Gordon, I. E., and Coauthors, 2022: The HITRAN2020 molecular spectroscopic database. Journal of Quantitative Spectroscopy and Radiative Transfer, 277, 107949, https://doi.org/10.1016/j.jqsrt.2021.107949.
    Guo, Y. X., and Coauthors, 2020: Air quality, nitrogen use efficiency and food security in China are improved by cost-effective agricultural nitrogen management. Nature Food, 1, 648−658, https://doi.org/10.1038/s43016-020-00162-z.
    Hao, L. Q., E. Kari, A. Leskinen, D. R. Worsnop, and A. Virtanen, 2020: Direct contribution of ammonia to α-pinene secondary organic aerosol formation. Atmospheric Chemistry and Physics, 20 , 14 393−14 405, https://doi.org/10.5194/acp-20-14393-2020.
    Hersbach, H., and Coauthors, 2020: The ERA5 global reanalysis. Quart. J. Roy. Meteor. Soc., 146, 1999−2049, https://doi.org/ 10.1002/qj.3803.
    Hilton, F., and Coauthors, 2012: Hyperspectral earth observation from IASI: Five years of accomplishments. Bull. Amer. Meteor. Soc., 93, 347−370, https://doi.org/10.1175/BAMS-D-11-00027.1.
    Inness, A., and Coauthors, 2019: The CAMS reanalysis of atmospheric composition. Atmospheric Chemistry and Physics, 19, 3515−3556, https://doi.org/10.5194/acp-19-3515-2019.
    Kaiser, J. W., and Coauthors, 2012: Biomass burning emissions estimated with a global fire assimilation system based on observed fire radiative power. Biogeosciences, 9, 527−554, https://doi.org/10.5194/bg-9-527-2012.
    Li, S. Q., H. Hu, C. G. G. Fang, S. C. Wang, S. P. Xun, B. F. He, W. Y. Wu, and Y. F. Huo, 2022: Hyperspectral infrared atmospheric sounder (HIRAS) atmospheric sounding system. Remote Sensing, 14, 3882, https://doi.org/10.3390/rs1416 3882.
    Luo, Z. Q., Y. Z. Zhang, W. Chen, M. Van Damme, P.-F. Coheur, and L. Clarisse, 2022: Estimating global ammonia (NH3) emissions based on IASI observations from 2008 to 2018. Atmospheric Chemistry and Physics, 22 , 10 375−10 388, https://doi.org/10.5194/acp-22-10375-2022.
    Mahowald, N. M., R. Scanza, J. Brahney, C. L. Goodale, P. G. Hess, J. K. Moore, and J. Neff, 2017: Aerosol deposition impacts on land and ocean carbon cycles. Current Climate Change Reports, 3, 16−31, https://doi.org/10.1007/s40641-017-0056-z.
    Mishchenko, M. I., and L. D. Travis, 1998: Capabilities and limitations of a current FORTRAN implementation of the T-matrix method for randomly oriented, rotationally symmetric scatterers. Journal of Quantitative Spectroscopy and Radiative Transfer, 60, 309−324, https://doi.org/10.1016/S0022-4073(98)00008-9.
    Na, K., C. Song, C. Switzer, and D. R. Cocker, 2007: Effect of ammonia on secondary organic aerosol formation from α-pinene ozonolysis in dry and humid conditions. Environ. Sci. Technol., 41, 6096−6102, https://doi.org/10.1021/es061956y.
    Qi, C. L., and Coauthors, 2020: High spectral infrared atmospheric sounder (HIRAS): System overview and on-orbit performance assessment. IEEE Trans. Geosci. Remote Sens., 58, 4335−4352, https://doi.org/10.1109/TGRS.2019.2963085.
    Rodgers, C. D., 2000: Inverse Methods for Atmospheric Sounding: Theory and Practice. World Scientific, 256 pp. https://doi.org/10.1142/3171.
    Shephard, M. W., and K. E. Cady-Pereira, 2015: Cross-track infrared sounder (CrIS) satellite observations of tropospheric ammonia. Atmospheric Measurement Techniques, 8, 1323−1336, https://doi.org/10.5194/amt-8-1323-2015.
    Siddans, R., D. Knappett, B. Kerridge, A. Waterfall, J. Hurley, B. Latter, H. Boesch, and R. Parker, 2017: Global height-resolved methane retrievals from the infrared atmospheric sounding interferometer (IASI) on MetOp. Atmospheric Measurement Techniques, 10, 4135−4164, https://doi.org/10.5194/amt-10-4135-2017.
    Sindelarova, K., and Coauthors, 2014: Global data set of biogenic VOC emissions calculated by the MEGAN model over the last 30 years. Atmospheric Chemistry and Physics, 14, 9317−9341, https://doi.org/10.5194/acp-14-9317-2014.
    Someya, Y., R. Imasu, K. Shiomi, and N. Saitoh, 2020: Atmospheric ammonia retrieval from the TANSO-FTS/GOSAT thermal infrared sounder. Atmospheric Measurement Techniques, 13, 309−321, https://doi.org/10.5194/amt-13-309-2020.
    Spurr, R., 2008: LIDORT and VLIDORT: Linearized pseudo-spherical scalar and vector discrete ordinate radiative transfer models for use in remote sensing retrieval problems. Light Scattering Reviews 3: Light Scattering and Reflection, A. A. Kokhanovsky, Ed., Springer, 229−275. https://doi.org/10.1007/978-3-540-48546-9_7.
    Stein, O., M. G. Schultz, I. Bouarar, H. Clark, V. Huijnen, A. Gaudel, M. George, and C. Clerbaux, 2014: On the wintertime low bias of Northern Hemisphere carbon monoxide found in global model simulations. Atmospheric Chemistry and Physics, 14, 9295−9316, https://doi.org/10.5194/acp-14-9295-2014.
    Twigg, M. M., and Coauthors, 2022: Intercomparison of in situ measurements of ambient NH3: Instrument performance and application under field conditions. Atmospheric Measurement Techniques, 15, 6755−6787, https://doi.org/10.5194/amt-15-6755-2022.
    Van Damme, M., J. W. Erisman, L. Clarisse, E. Dammers, S. Whitburn, C. Clerbaux, A. J. Dolman, and P.-F. Coheur, 2015: Worldwide spatiotemporal atmospheric ammonia (NH3) columns variability revealed by satellite. Geophys. Res. Lett., 42, 8660−8668, https://doi.org/10.1002/2015GL0 65496.
    Van Damme, M., L. Clarisse, S. Whitburn, J. Hadji-Lazaro, D. Hurtmans, C. Clerbaux, and P.-F. Coheur, 2018: Industrial and agricultural ammonia point sources exposed. Nature, 564, 99−103, https://doi.org/10.1038/s41586-018-0747-1.
    Vandaele, A. C., M. Kruglanski, and M. De Mazière, 2006: Modeling and retrieval of atmospheric spectra using ASIMUT. Proc. of the First “Atmospheric Science Conference”, Vol. ESA SP-628, July 2006, Frascati, Italy.
    Vandenbussche, S., S. Kochenova, A. C. Vandaele, N. Kumps, and M. De Mazière, 2013: Retrieval of desert dust aerosol vertical profiles from IASI measurements in the TIR atmospheric window. Atmospheric Measurement Techniques, 6, 2577−2591, https://doi.org/10.5194/amt-6-2577-2013.
    Waliser, D. E., and C. Gautier, 1993: A satellite-derived climatology of the ITCZ. J. Climate, 6, 2162−2174, https://doi.org/10.1175/1520-0442(1993)006<2162:ASDCOT>2.0.CO;2.
    Wang, R., and Coauthors, 2021: Monthly patterns of ammonia over the contiguous United States at 2‐km resolution. Geophys. Res. Lett., 48, e2020GL090579, https://doi.org/10.1029/2020GL090579.
    Whitburn, S., L. Clarisse, M. Crapeau, T. August, T. Hultberg, P. F. Coheur, and C. Clerbaux, 2022: A CO2-independent cloud mask from Infrared atmospheric sounding interferometer (IASI) radiances for climate applications. Atmospheric Measurement Techniques, 15, 6653−6668, https://doi.org/10.5194/amt-15-6653-2022.
    Wu, C. Q., and Coauthors, 2020: FY-3D HIRAS radiometric calibration and accuracy assessment. IEEE Trans. Geosci. Remote Sens., 58, 3965−3976, https://doi.org/10.1109/TGRS.2019.2959830.
    Xian, D., P. Zhang, L. Gao, R. J. Sun, H. Z. Zhang, and X. Jia, 2021: Fengyun meteorological satellite products for earth system science applications. Adv. Atmos. Sci., 38, 1267−1284, https://doi.org/10.1007/s00376-021-0425-3.
    Xu, L., and J. E. Penner, 2012: Global simulations of nitrate and ammonium aerosols and their radiative effects. Atmospheric Chemistry and Physics, 12, 9479−9504, https://doi.org/10.5194/acp-12-9479-2012.
    Yang, Z. D., and Coauthors, 2019: Capability of Fengyun-3D satellite in earth system observation. J. Meteor. Res., 33, 1113−1130, https://doi.org/10.1007/s13351-019-9063-4.
    Zhang, C. M., M. J. Gu, Y. Hu, P. Y. Huang, T. H. Yang, S. Huang, C. L. Yang, and C. Y. Shao, 2021: A study on the retrieval of temperature and humidity profiles based on FY-3D/HIRAS infrared hyperspectral data. Remote Sensing, 13, 2157, https://doi.org/10.3390/rs13112157.
    Zhou, D. K., A. M. Larar, X. Liu, W. L. Smith, L. L. Strow, P. Yang, P. Schlüssel, and X. Calbet, 2011: Global land surface emissivity retrieved from satellite ultraspectral IR measurements. IEEE Trans. Geosci. Remote Sens., 49, 1277−1290, https://doi.org/10.1109/TGRS.2010.2051036.
    Zhu, L. Y., D. K. Henze, J. O. Bash, K. E. Cady-Pereira, M. W. Shephard, M. Luo, and S. L. Capps, 2015: Sources and impacts of atmospheric NH3: Current understanding and frontiers for modeling, measurements, and remote sensing in North America. Current Pollution Reports, 1, 95−116, https://doi.org/10.1007/s40726-015-0010-4.
  • [1] WANG Hesong, JIA Gensuo, 2013: Regional Estimates of Evapotranspiration over Northern China Using a Remote-sensing-based Triangle Interpolation Method, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 1479-1490.  doi: 10.1007/s00376-013-2294-x
    [2] Zhao Gaoxiang, 1998: Analysis of the Ability of Infrared Water Vapor Channel for Moisture Remote Sensing in the Lower Atmosphere, ADVANCES IN ATMOSPHERIC SCIENCES, 15, 107-112.  doi: 10.1007/s00376-998-0022-8
    [3] Yunji ZHANG, Eugene E. CLOTHIAUX, David J. STENSRUD, 2022: Correlation Structures between Satellite All-Sky Infrared Brightness Temperatures and the Atmospheric State at Storm Scales, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 714-732.  doi: 10.1007/s00376-021-0352-3
    [4] Ping YANG, Kuo-Nan LIOU, Lei BI, Chao LIU, Bingqi YI, Bryan A. BAUM, 2015: On the Radiative Properties of Ice Clouds: Light Scattering, Remote Sensing, and Radiation Parameterization, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 32-63.  doi: 10.1007/s00376-014-0011-z
    [5] QIU Jinhuan, CHEN Hongbin, 2004: Recent Progresses in Atmospheric Remote Sensing Research in China-- Chinese National Report on Atmospheric Remote Sensing Research in China during 1999-2003, ADVANCES IN ATMOSPHERIC SCIENCES, 21, 475-484.  doi: 10.1007/BF02915574
    [6] Xiangqian WU, Changyong CAO, 2006: Sensor Calibration in Support for NOAA’s Satellite Mission, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 80-90.  doi: 10.1007/s00376-006-0009-2
    [7] WANG Hesong, JIA Gensuo, 2012: Satellite-Based Monitoring of Decadal Soil Salinization and Climate Effects in a Semi-arid Region of China, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 1089-1099.  doi: 10.1007/s00376-012-1150-8
    [8] Zhonghui Tan, ju wang, Jianping Guo, Chao Liu, miao zhang, Shuo Ma, 2024: Improving satellite-retrieved cloud base height with ground-based cloud radar measurements, ADVANCES IN ATMOSPHERIC SCIENCES.  doi: 10.1007/s00376-024-4052-7
    [9] Minqiang ZHOU, Qichen NI, Zhaonan CAI, Bavo LANGEROCK, Jingyi JIANG, Ke CHE, Jiaxin WANG, Weidong NAN, Yi LIU, Pucai WANG, 2023: Ground-Based Atmospheric CO2, CH4, and CO Column Measurements at Golmud in the Qinghai-Tibetan Plateau and Comparisons with TROPOMI/S5P Satellite Observations, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 223-234.  doi: 10.1007/s00376-022-2116-0
    [10] Xiaoxiong XIONG, William BARNES, 2006: An Overview of MODIS Radiometric Calibration and Characterization, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 69-79.  doi: 10.1007/s00376-006-0008-3
    [11] Lu YAO, Dongxu YANG, Yi LIU, Jing WANG, Liangyun LIU, Shanshan DU, Zhaonan CAI, Naimeng LU, Daren LYU, Maohua WANG, Zengshan YIN, Yuquan ZHENG, 2021: A New Global Solar-induced Chlorophyll Fluorescence (SIF) Data Product from TanSat Measurements, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 341-345.  doi: 10.1007/s00376-020-0204-6
    [12] Dongxu YANG, Yi LIU, Hartmut BOESCH, Lu YAO, Antonio DI NOIA, Zhaonan CAI, Naimeng LU, Daren LYU, Maohua WANG, Jing WANG, Zengshan YIN, Yuquan ZHENG, 2021: A New TanSat XCO2 Global Product towards Climate Studies, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 8-11.  doi: 10.1007/s00376-020-0297-y
    [13] HE Yuting, JIA Gensuo, HU Yonghong, and ZHOU Zijiang, 2013: Detecting urban warming signals in climate records, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 1143-1153.  doi: 10.1007/s00376-012-2135-3
    [14] Ke CHE, Yi LIU, Zhaonan CAI, Dongxu YANG, Haibo WANG, Denghui JI, Yang YANG, Pucai WANG, 2022: Characterization of Regional Combustion Efficiency using ΔXCO: ΔXCO2 Observed by a Portable Fourier-Transform Spectrometer at an Urban Site in Beijing, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 1299-1315.  doi: 10.1007/s00376-022-1247-7
    [15] Dongxu YANG, Janne HAKKARAINEN, Yi LIU, Iolanda IALONGO, Zhaonan CAI, Johanna TAMMINEN, 2023: Detection of Anthropogenic CO2 Emission Signatures with TanSat CO2 and with Copernicus Sentinel-5 Precursor (S5P) NO2 Measurements: First Results, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 1-5.  doi: 10.1007/s00376-022-2237-5
    [16] ZHOU Suoquan, CHEN Jingming, GONG Peng, XUE Genyuan, 2006: Effects of Heterogeneous Vegetation on the Surface Hydrological Cycle, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 391-404.  doi: 10.1007/s00376-006-0391-9
    [17] WANG Dongxiao, ZHANG Yan, ZENG Lili, LUO Lin, 2009: Marine Meteorology Research Progress of China from 2003 to 2006, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 17-30.  doi: 10.1007/s00376-009-0017-0
    [18] BAO Qing, LIU Yimin, SHI Jiancheng, WU Guoxiong, 2010: Comparisons of Soil Moisture Datasets over the Tibetan Plateau and Application to the Simulation of Asia Summer Monsoon Onset, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 303-314.  doi: 10.1007/s00376-009-8132-5
    [19] Fuzhong WENG, Xinwen YU, Yihong DUAN, Jun YANG, Jianjie WANG, 2020: Advanced Radiative Transfer Modeling System (ARMS): A New-Generation Satellite Observation Operator Developed for Numerical Weather Prediction and Remote Sensing Applications, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 131-136.  doi: 10.1007/s00376-019-9170-2
    [20] Cheng Minghu, Shi Guangyu, Zhou Xiuji, 1990: Numerical Experiment of Combined Infrared and Ultraviolet Radiation Remote Sensing to Determine the Profile and Total Content of Atmospheric Ozone, ADVANCES IN ATMOSPHERIC SCIENCES, 7, 305-319.  doi: 10.1007/BF03179763

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 21 March 2023
Manuscript revised: 13 July 2023
Manuscript accepted: 14 July 2023
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

The First Global Map of Atmospheric Ammonia (NH3) as Observed by the HIRAS/FY-3D Satellite

    Corresponding author: Xingying ZHANG, zxy@cma.gov.cn
  • 1. Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing 100029, China
  • 2. Royal Belgian Institute for Space Aeronomy (BIRA-IASB), Brussels 1180, Belgium
  • 3. National Satellite Meteorology Center, China Meteorological Administration, Beijing 100081, China
  • 4. Department of Science & Technology and Climate Change, China Meteorological Administration, Beijing 100081, China
  • 5. University of Chinese Academy of Sciences, Beijing 100049, China

Abstract: Atmospheric ammonia (NH3) is a chemically active trace gas that plays an important role in the atmospheric environment and climate change. Satellite remote sensing is a powerful technique to monitor NH3 concentration based on the absorption lines of NH3 in the thermal infrared region. In this study, we establish a retrieval algorithm to derive the NH3 column from the Hyperspectral Infrared Atmospheric Sounder (HIRAS) onboard the Chinese FengYun (FY)-3D satellite and present the first atmospheric NH3 column global map observed by the HIRAS instrument. The HIRAS observations can well capture NH3 hotspots around the world, e.g., India, West Africa, and East China, where large NH3 emissions exist. The HIRAS NH3 columns are also compared to the space-based Infrared Atmospheric Sounding Interferometer (IASI) measurements, and we find that the two instruments observe a consistent NH3 global distribution, with correlation coefficient (R) values of 0.28–0.73. Finally, some remaining issues about the HIRAS NH3 retrieval are discussed.

摘要: 氨气(NH3)在地球大气中是一种化学性质活泼的微量气体,在大气环境和气候变化中发挥着重要作用。卫星遥感是借助于NH3在热红外谱段的特征吸收线,反演获得NH3浓度的有效技术。在本研究中,基于最优估计理论建立了一套NH3柱浓度的反演算法,成功应用于中国风云-3D卫星上的高光谱红外大气探测仪(HIRAS),获得了HIRAS仪器观测到的大气NH3柱全球分布图。风云-3D卫星观测可以很好地捕捉全球NH3高值区,例如印度、西非、中国东部等存在大量NH3排放的地区。风云-3D卫星与欧洲红外大气探测干涉仪卫星(IASI)测量的NH3结果具有良好的一致性,相关系数 (R) 值为0.28–0.73。最后,本文针对风云-3D/HIRAS卫星NH3反演目前仍存在一些问题进行讨论,并提出了未来风云卫星发展方向。

    • Atmospheric ammonia (NH3) is an important trace gas that causes environmental problems and harms human health. As the most abundant alkaline gas, NH3 reacts with atmospheric sulfuric acid and nitric acid, and it is a major source of aerosols (Na et al., 2007; Hao et al., 2020; Bao et al., 2021). The ammonium aerosols change the reflectivity of the Earth’s surface and play a significant role in the radiance balance of the Earth’s system (Xu and Penner, 2012). When aerosols are deposited on water surfaces, eutrophication can occur, damaging the aquatic ecosystem (Mahowald et al., 2017). Guo et al. (2020) estimated that ammonia-related aerosols could lead to 36 billion dollars of public health loss per year in the United States.

      Ammonia (NH3) in the atmosphere is mainly emitted from livestock animal wastes and fertilizer productions/applications (~85%), and partly from industrial manufacture, biomass burning, transports, plant decomposition, and volatilization from oceans and soils (Behera et al., 2013; Zhu et al., 2015; Van Damme et al., 2018). Due to its importance, NH3 emission inventories have been developed (Bouwman et al., 1997; Crippa et al., 2020). However, the current NH3 emissions remain highly uncertain at a global scale due to a lack of NH3 observations (Luo et al., 2022). It is quite difficult to measure the NH3 concentration by in situ techniques, as the NH3 molecules will adsorb to the inlet and other surfaces of the in situ instrument (Twigg et al., 2022).

      Satellite remote sensing techniques using the specific absorption lines of NH3 in the thermal infrared spectral region can provide global NH3 column measurements. The first global NH3 map was successfully derived from the Infrared Atmospheric Sounding Interferometer (IASI) carried on the MetOp-A satellite (Clarisse et al., 2009). Later, NH3 columns were also derived from several other hyperspectral infrared satellites, e.g., the Cross-track Infrared Sounder (CrIS) and the Thermal And Near infrared Sensor for carbon Observations-Fourier Transform Spectrometer (TANSO-FTS) (Shephard and Cady-Pereira, 2015; Someya et al., 2020). The NH3 columns derived from satellite IASI and CrIS measurements have been compared and validated with ground-based Fourier Transform Spectrometer (FTS) retrievals, and the results show that there is a good correlation between ground-based and satellite NH3 measurements and the bias between them is typically within their estimated retrieval uncertainties (Dammers et al., 2016, 2017).

      In November 2017, the Hyperspectral Infrared Atmospheric Sounder (HIRAS) was launched into a sun-synchronous polar orbit at 863 km above the ground onboard the meteorological FengYun (FY)-3D satellite. It was the first Chinese satellite sensor to provide nadir hyperspectral infrared spectra of the upwelling radiance between 650 and 2550 cm–1 (Yang et al., 2019). The performance of HIRAS has been well evaluated and calibrated in the optical laboratory before launch, and in space after launch (Qi et al., 2020; Wu et al., 2020). The brightness temperature noise level of the HIRAS spectra is about 0.15–0.4 K at 280 K (Qi et al., 2020), which is slightly larger than the IASI noise level of about 0.1–0.2 K at 280 K (Hilton et al., 2012). Atmospheric temperature and water vapor have been successfully derived from the HIRAS spectra (Zhang et al., 2021; Li et al., 2022). However, until now, the HIRAS spectra have never been used to retrieve atmospheric NH3 concentrations.

      In this study, we establish a retrieval algorithm to derive the NH3 column from the HIRAS observed spectra and present the first HIRAS NH3 global map. Section 2 describes the satellite data and the retrieval method. In section 3, we show the retrieval results and the comparisons between HIRAS and IASI NH3 column retrievals. After that, the current issues affecting HIRAS NH3 retrieval are discussed in section 4. Finally, conclusions are drawn in section 5.

    2.   Data and retrieval methods
    • HIRAS is a Fourier transform spectrometer that records infrared spectra emitted by the Earth, its atmosphere, and the solar component. One HIRAS cross-track scan sequence has 33 measurements, including 29 Earth Scenes, 2 Deep Space, and 2 Internal Calibration Target observations (Wu et al., 2020). With a wide swath width of about 2250 km, it offers near-global coverage twice a day, with overpass times at 0130 and 1330 (local solar time). The field of view (FOV) of each HIRAS footprint is 1.1°, corresponding to a 16-km diameter projection on the ground at nadir (Qi et al., 2020). There are three bands observed by the HIRAS sensor: the long wave IR (LWIR) band from 650 to 1135 cm–1, the middle wave IR (MWIR) band from 1210 to 1750 cm–1, and the short wave IR (SWIR) band from 2155 to 2550 cm–1. All bands are recorded at the same full spectral resolution of 0.625 cm–1, corresponding to a maximum optical path difference (MOPD) of 0.8 cm. For the MWIR and SWIR bands, HIRAS also provides data with spectral resolutions of 1.25 and 2.5 cm–1, respectively. As the NH3 absorption lines lie mainly in the LWIR (Gordon et al., 2022), only the HIRAS LWIR spectra with a spectral resolution of 0.625 cm–1 are used hereafter.

      Figure 1a shows a typical HIRAS LWIR spectrum, along with the NH3 retrieval window. The HIRAS spectra used in this study cover January and July 2020 (2 months; 62 days). According to the HITRAN 2020 spectroscopy, relatively strong NH3 absorption lines are found near 930 and 967 cm–1. We tested both the 925–935 cm–1 and 960–970 cm–1 spectral windows to check their sensitivity to the NH3 columns. In the former window, we found that there are many strong CO2 and H2O lines that contaminate the NH3 signal. Therefore, the latter (960–970 cm–1) window was selected for retrieval. Figure 2 shows a typical transmittance spectrum from the surface to the top of the atmosphere between 960 and 970 cm–1 using the US standard atmosphere (NOAA, 1976). Although there is still some interference from CO2, H2O, and O3; NH3 absorption lines are less affected by them, especially near 965.5 and 967.3 cm–1.

      Figure 1.  (a) A typical HIRAS observed spectra in the long wavelength range (648–1136 cm–1; spectral resolution of 0.625 cm–1) at 18.72°N, 70.50°E, for a satellite zenith angle of 3.27° and a solar zenith angle of 45.25°. The NH3 retrieval window (960–970 cm–1) is marked by a red shadow. (b) The observed (Obs) and simulated (Sim) spectra, and (c) the residual of the fitted spectrum (Obs – Sim).

      Figure 2.  (a) The transmittances of the main species (CO2, H2O, O3, and NH3) in the NH3 retrieval window, and (b) a zoom window in the transmittance range between 0.965 and 1.00 to establish a better view of NH3 absorption lines.

      To reduce the impact from clouds, we only perform the NH3 retrieval under clear-sky conditions. The medium-resolution spectral imager-2 (MERSI-2) sensor onboard the FY-3D satellite provides cloud mask products with a 250-m spatial resolution (Xian et al., 2021). We calculate the cloud fraction based on the MERSI-2 cloud measurements for each HIRAS observation and select the clear-sky HIRAS measurements (cloud fraction equal to 0). The cloud masking procedure used here is the same as in Li et al. (2022).

    • The optimal estimation method (OEM; Rodgers, 2000) is applied to retrieve the NH3 column from the observed HIRAS spectra. A cost function [$ \mathrm{J}\left(\mathbf{x}\right) $] is defined by Eq. (1):

      where $ \mathbf{y} $ is the observed spectra; $ \mathbf{F}\left(\mathbf{x},\mathbf{b}\right) $ is the forward model to simulate spectra; $ \mathbf{x} $ is the state vector (retrieved parameters); $ \mathbf{b} $ represents the model parameters that are not retrieved, such as the surface emissivity, temperature profile, and satellite-Earth geometry; subscripts $ \mathrm{a} $ and $ \epsilon $ represent the prior and the measurement noise, respectively; $ {\mathbf{x}}_{\mathrm{a}} $ is the prior, presenting the best estimation of the state vector based on the a priori knowledge; $ {\mathbf{S}}_{\epsilon } $ is the measurement covariance matrix, determined by the signal-to-noise-ratio (SNR) of the HIRAS observed spectrum (the diagonal values of the $ {\mathbf{S}}_{\epsilon } $ are calculated as $ 1/{\mathrm{S}\mathrm{N}\mathrm{R}}^{2} $, and the non-diagonal values are set to 0); $ {\mathbf{S}}_{\mathrm{a}} $ is the a priori covariance matrix of the state vector, derived from an atmospheric chemistry transport model. The Newton iteration is applied to find the approximation of the true state which agrees best with both the measurement and the a priori information. It follows:

      where G is the contribution matrix, K is the Jacobian matrix, representing the sensitivity of the observed spectra to the parameters, and subscript i is the iteration index. Finally, the optimal state vector ($ \widehat{\mathbf{x}} $) is given by Eq. (4):

      where $ \mathbf{A} $ is the averaging kernel matrix, representing the sensitivity of the retrieved parameters to the true state; $ \epsilon $ is the retrieval uncertainty.

      Table 1 lists the parameters in the state vector ($ \mathbf{x} $), together with both their a priori and variance settings. Apart from the NH3 column, the columns of the interfering species (CO2, H2O, and O3) are retrieved as well. In addition, the spectral shift and the surface temperature are included. Since the spatiotemporal variations of the H2O column and the surface temperature are very large, their a priori values are derived from the ERA5 hourly reanalysis data (Hersbach et al., 2020).

      Retrieved parameters ($ \mathbf{x} $) Prior 1σ
      NH3 column CAMS model SD of CAMS monthly means
      O3 column CAMS model 50%
      CO2 column Carbon Tracker 10%
      H2O column ERA5 reanalysis data 100%
      Surface temperature ERA5 reanalysis data 3%
      Spectral shift 0 10%

      Table 1.  The state vector ($ \mathbf{x} $) together with their prior and variance (1σ) in the HIRAS NH3 retrieval algorithm.

    • We use the ASIMUT model to simulate the infrared radiance transmitted from the Earth’s surface to the HIRAS satellite sensor. The ASIMUT is a radiative transfer model (Vandaele et al., 2006) developed by the Royal Belgian Institute for Space Aeronomy (BIRA-IASB) which calculates the spectrum and analytical derivation of the Jacobians. The ASIMUT code has been applied for the dust and methane retrievals from the IASI satellite (Vandenbussche et al., 2013; De Wachter et al., 2017). The ASIMUT software has been coupled to the SPHER/TMATRIX (Mishchenko and Travis, 1998) and (V)LIDORT (Spurr, 2008) to compute the atmospheric scatterings.

      In the NH3 retrieval window (960–970 cm–1), we neglect the scattering and consider the thermal emissions under local thermodynamical equilibrium. The line-by-line (LBL) method to calculate the cross-sections was implemented in the ASIMUT model, but it consumes large computing resources. To speed up the retrieval, we create look-up-tables (LUTs) for CO2, H2O, O3, and NH3 cross-sections in the spectral range from 900 to 1000 cm–1 based on the HITRAN 2020 database (Gordon et al., 2022), for various pressures (between 7 × 10–4 hPa and 1081 hPa) and temperatures (between 148 and 328 K). Numerous simulations with different pressure and temperature conditions have been carried out with both the LBL method and the LUT method, and the relative radiance differences between the LBL and LUT simulations are all within 0.005%, which is much smaller than the noise level of the HIRAS spectra.

      The ASIMUT model includes 39 vertical layers between the surface and the top of the atmosphere (40 vertical levels: 0.0, 0.1, 0.3, 0.5, 0.7, 0.9, 1.0, 1.3, 1.5, 1.75, 2.0, 2.5, 3.2, 3.75, 4.5, 5.0, 6.0, 7.0, 8.0, 9.0, 10.0, 11.0, 12.0, 13.0, 14.0, 15.0, 17.0, 19.0, 21.0, 23.0, 25.0, 27.0, 30.0, 35.0, 40.0, 45.0, 50.0, 60.0, 70.0, and 100.0 km). The global spectrally dependent surface emissivity datasets are provided by Zhou et al. (2011). The HIRAS instrument line shape (ILS) has been taken into account in the forward model, which is characterized by a sinc function with a 0.625 cm–1 spectral resolution.

    • The Copernicus Atmosphere Monitoring Service (CAMS) global ECMWF Atmospheric Composition Reanalysis 4 (EAC4) model simulations are applied to generate the a priori information of NH3. The CAMS EAC4 model has a horizontal resolution of 0.75° × 0.75°, with 60 model vertical levels between the ground and 0.2 hPa. The CAMS model uses anthropogenic emissions from the MACCity inventory (Stein et al., 2014), fire emissions from the Global Fire Assimilation System (GFAS) (Kaiser et al., 2012), and the Model of Emissions of Gases and Aerosols from Nature (MEGAN) driven by the MERRA reanalyzed meteorology to generate the monthly mean volatile organic compound emissions (Sindelarova et al., 2014). For more information about the CAMS model simulations refer to Inness et al. (2019) and the references therein. Previous studies demonstrate that there are large day-to-day and month-to-month variabilities of NH3 globally (Van Damme et al., 2015; Wang et al., 2021). Therefore, for the HIRAS NH3 retrieval, we use the CAMS model monthly means between 2015 and 2020 (6 years) to generate the a priori profile of NH3. Moreover, the standard deviation (std) of the daily NH3 concentration is calculated in each grid cell to set the variability of the NH3 (Sa = 1$ {\sigma }^{2} $). Figure 3 shows the CAMS simulated NH3 mole fraction near the surface (model bottom level). The NH3 mole fractions are relatively high in Southeast Asia, South Asia, Europe, Australia, North America, and South America, and the NH3 mole fractions are relatively low above the ocean and in the polar region, which is generally consistent with the IASI satellite observations (Van Damme et al., 2015). Figure 4a shows typical NH3 vertical profiles over land and ocean derived from the CAMS model. The NH3 mole fraction generally decreases with altitude and becomes less than 0.01 ppb above 10 km.

      Figure 3.  The CAMS model global simulated NH3 mole fraction near the surface in January, April, July, and October between 2015 and 2020.

      Figure 4.  Typical NH3 vertical profiles from the CAMS model over land and ocean. (a) Typical averaging kernel matrix (solid lines; lines are colored with different altitudes; in units of ppb (ppb)−1 and column averaging kernel scaled with 0.1 (red dashed line; in units of (molecules cm–2) (molecules cm–2)−1 of the HIRAS NH3 retrieval over (b) land and (c) ocean.

    • As shown in Table 1, we perform the column retrieval (only a scaling factor of NH3 column in the state vector) for the HIRAS NH3 retrieval. To understand the vertical sensitivity of the NH3 retrieved column, we carried out profile retrieval for NH3 above several areas, such as in India and the tropical ocean, to derive the NH3 profile averaging kernel (AVK) matrix in units of ppb (ppb)−1. Note that we have tuned the Sa values to make the sum of the AVK matrix (DOF) from the profile retrieval close to what we get from the column retrieval. Then the column-averaging kernel (CAVK) is calculated based on the AVK matrix as follows:

      where $ {\mathbf{x}}_{\mathrm{a},\mathrm{p}} $, $ {\mathbf{x}}_{\mathrm{t},\mathrm{p}} $, and $ {\mathbf{x}}_{\mathrm{r},\mathrm{p}} $ are the a priori, true, and retrieved NH3 vertical mole fraction profiles, respectively; $ \mathrm{T}{\mathrm{C}}_{\mathrm{a}} $ and $ \mathrm{T}{\mathrm{C}}_{\mathrm{r}} $ are the a priori and retrieved NH3 column, respectively, in units of molecules cm–2; $ {\bf{PC}}_{\mathrm{a}} $ and $ \bf{P}{\bf{C}}_{\mathrm{t}} $ are the a priori and true partial column vertical profiles of NH3, respectively, in a unit of molecules cm–2; $ \bf{C}\bf{A}\bf{V}\bf{K} $ is the column averaging kernel in units of (molecules cm–2) (molecules cm–2)–1; $ {\bf{P}\bf{C}}_{\mathrm{a}\mathrm{i}\mathrm{r}} $ is the partical column vertical profile of the dry air.

      Figures 4b and 4c show a typical averaging kernel (AVK) of HIRAS retrieved NH3 over land and ocean. Due to the weak absorption lines and low concentrations of NH3, the retrieved HIRAS NH3 column is mainly sensitive to the mid-troposphere (2–8 km), which is similar to GOSAT and TES satellite retrievals (Clarisse et al., 2010; Someya et al., 2020). The NH3 mole fraction is high in the boundary layer (0–2 km), but it also shows relatively high values in the mid-troposphere (2–6 km; Fig. 4a), especially over land. Consequently, such an AVK still allows us to derive NH3 column information. Note that the AVK value varies strongly with surface type and atmospheric conditions. For instance, the NH3 signals captured by HIRAS over the ocean are very weak leading to a degree of freedom (DOFs) close to 0, and the DOFs can be up to 0.9 over several polluted land areas.

    3.   Results
    • As an example, Fig. 5 shows the NH3 global maps observed by the HIRAS satellite during daytime and nighttime on 30 January 2020. Note that we filter out the retrieved NH3 columns with a DOF < 0.2, root mean square error of the fitting residual (RMSE) > 0.55%, and an NH3 column < 0. More valid NH3 column data are available over land than over the ocean. Moreover, the uncertainty of the NH3 retrievals is relatively large during the nighttime as compared to the daytime, because the NH3 retrieval quality depends on the thermal contrast between the surface and the lowest atmospheric layer; therefore, it is more challenging to capture the NH3 signal near the surface when the thermal contrast is small during nighttime (Clarisse et al., 2010).

      Figure 5.  The retrieved NH3 columns on 30 January 2020 from the HIRAS observations under clear-sky conditions during daytime (a) and nighttime (c), together with the differences between the retrieved columns and a priori columns (re-ap) during daytime (b) and nighttime (d), respectively.

      The HIRAS measurements show several NH3 hotspots around the world, e.g., India, West Africa, and East China. These hotspots observed in the HIRAS data are generally consistent with the CAMS model simulations (Fig. 3) since there are large NH3 emissions in these regions (Stein et al., 2014). Compared to the CAMS model monthly means between 2015 and 2020, the HIRAS retrieved NH3 columns during the daytime on 30 January 2020 are larger than the model simulations in middle Africa and lower than the model simulations in East China.

    • To assess the uncertainty of the HIRAS NH3 retrievals, we compare HIRAS with IASI/Metop-B NH3 measurements. The IASI measurements have been validated with nine ground-based FTIR sites around the world, and the relative means and stds of the differences (IASI – FTIR) are –32.4% ±56.3% (Dammers et al., 2016). Figures 6 and 7 show the global monthly NH3 columns during daytime observed by HIRAS and IASI satellites in January and July 2020, respectively. The HIRAS and IASI data pairs are selected based on their co-located gridded monthly values. Although HIRAS provides lower data density than IASI, especially over the ocean and in the high-latitude regions, the two satellites see a similar global map of NH3 columns. High NH3 columns are observed in regions with high human activities or strong biomass burnings, which agrees well with bottom-up emission inventories (Crippa et al., 2020).

      Figure 6.  The NH3 column monthly means in January 2020 observed by (a) HIRAS/FY3D and (b) IASI/MetopB re-gridded onto 1° × 1° (latitude × longitude). The scatter plots between HIRAS and IASI NH3 columns in (c) North America, (d) Europe, (e) East China, (f) South America, (g) Africa, and (h) India. In each scatter plot, the dots are colored according to the data density. The red dashed line is the linear regression. The solid black line is the one-to-one line. R is the Pearson correlation coefficient and N is the number of the data points.

      Figure 7.  Same as Fig. 6, but for July 2020.

      Six regions (North America, Europe, East China, South America, Africa, and India) are selected to investigate the correlations and relative differences between HIRAS and IASI NH3 monthly mean columns (Table 2). Positive correlations between both satellite data sets are identified in all these regions, with the Pearson correlation coefficient (R) ranging from 0.28 to 0.73 in January, and from 0.31 to 0.58 in July. The R values are relatively lower in July as compared to those in January, especially in East China, Africa, and India, because the HIRAS has significantly fewer clean-sky pixels in these regions after the cloud filtering. Table 2 shows the means and standard deviations (stds) of the differences between HIRAS and IASI NH3 columns. In January, the mean relative differences (HIRAS-IASI) are within 5.5% in North America, Europe, Africa, and India, –17.2% in East China, and 35.6% in South America. The stds of their relative differences are between 46.6% and 82.4%. In July, the mean relative differences (HIRAS – IASI) are within 12% in Europe, East China, Africa, India, and 42.5% in South America. The stds of their relative differences are between 51.8% and 82.3%. These values are comparable to the means and stds of the differences between IASI and ground-based FTIR measurements.

      Region (latitude/longitude)
      North America Europe East China South America Africa India
      [15°N, 55°N]/
      [120°W, 80°W]
      [30°N, 60°N]/
      [20°W, 30°E]
      [15°N, 45°N]/
      [100°E, 120°E]
      [60°S, 0°]/
      [80°W, 30°W]
      [0°, 18°N]/
      [20°W, 60°E]
      [10°N, 35°N]/
      [60°E, 90°E]
      Jan-2020 Mean (%) –2.0 –3.1 –17.2 35.6 –5.5 –5.2
      Std (%) 52.9 65.1 65.1 82.4 51.7 46.6
      R 0.28 0.31 0.69 0.55 0.68 0.73
      P-value <0.001 <0.001 <0.001 <0.001 <0.001 <0.001
      DOF 0.47 0.48 0.61 0.53 0.49 0.57
      Jul-2020 Mean (%) 42.5 11.4 –3.0 –9.9 –11.0 –3.6
      Std (%) 78.7 53.4 82.3 71.5 51.8 60.8
      R 0.50 0.32 0.31 0.32 0.47 0.58
      P-value <0.001 <0.001 <0.001 <0.001 <0.001 <0.001
      DOF 0.57 0.66 0.44 0.45 0.42 0.54

      Table 2.  The relative mean and std of the difference between HIRAS and IASI measurements [(HIRAS−IASI)/IASI × 100%] over six regions with high NH3 columns in January and July 2020, together with their Pearson correlation coefficients (R), p-values, and mean DOFs.

    4.   Discussion
    • Currently, the HIRAS satellite after quality filtering provides a smaller data density than the IASI, and the HIRAS NH3 measurements are mainly over land and in mid- and low-latitude regions. Figure 8 shows all the convergent retrievals derived from the HIRAS spectra in January 2020 before applying the RMSE and DOF filtering. We found a strong latitudinal dependence in terms of the RMSE, with relatively low values in low-latitude regions and high values in high-latitude regions. Figure 9 illustrates the relationship between the RMSE and DOF of HIRAS NH3 retrievals in January 2020 and latitude. Similar to Fig. 8, we find that the RMSE is pretty large and the DOF is generally less than 0.05 in the high-latitude regions. The SNR of HIRAS spectra is low in high-latitude regions, which is probably caused by low surface temperatures and significant snow coverage. With our quality filtering criteria, the HIRAS NH3 retrievals in the high-latitude regions are mainly filtered out. In Fig. 8b, several areas in the tropics and near the Inter-tropical Convergence Zone (ITCZ; Waliser and Gautier, 1993) still have a relatively high RMSE which indicates remaining cloud effects. Therefore, an improved cloud screening method is needed in the future to better select clear-sky HIRAS measurements. For example, we could apply the nitrous oxide (N2O) infrared absorption lines to retrieve cloud parameters and to do a cloud mask (Siddans et al., 2017), or design a new cloud detection algorithm based on a neural network (Whitburn et al., 2022). Apart from the ITCZ regions, there is no obvious RMSE difference between land and ocean retrievals in tropical areas. However, we find that the DOF over the ocean is quite low even with a low RMSE. The reason is that the NH3 concentration over the ocean is so low that the NH3 signal is too weak to be captured from the HIRAS observed spectra. Regarding the HIRAS measurements in the high-latitude regions, the RMSE is generally larger than 0.5%, leading to a low DOF as well. Overall, for all the retrievals with a low DOF due to a bad fitting or a low NH3 column, the retrieved NH3 columns are pretty close to their a priori values as the retrieval is dominated by the a priori information (Rodgers, 2000).

      Figure 8.  (a) The NH3 columns, (b), the RMSE, and (c) the DOF for all convergent retrievals derived from the HIRAS spectra in January 2020.

      Figure 9.  The RMSE (left) and DOF (right) as a function of the latitude bin (every 10 degrees) as derived from all the HIRAS NH3 retrievals (grey) and the retrievals after quality filtering (orange) in January 2020. The error bar is the standard deviation.

      To better understand why there are almost no HIRAS NH3 retrievals left over the ocean after quality filtering, we have applied our retrieval algorithm developed for HIRAS (section 2) to IASI L1C data. Figure 10 shows typical spectra over the ocean derived from two adjacent HIRAS and IASI observations on 30 January 2020, respectively. Note that the same a priori profiles coming from the CAMS model are used for the HIRAS and IASI NH3 retrievals. Both retrievals are convergent. However, the DOF of the retrieved NH3 column from the HIRAS spectrum is very low (DOF = 0.03) causing the retrieved NH3 column to be almost the same as the a priori column. The RMSE of the IASI residual is 0.03%, which is about eight times better than that of the HIRAS retrieval (0.25%). Thanks to the high SNR, we can successfully retrieve the NH3 column (DOF = 0.78) from the IASI spectrum. This experiment demonstrates that the SNR of the HIRAS spectrum is currently too weak to derive the NH3 column above regions with a low NH3 column, such as the ocean.

      Figure 10.  Typical NH3 retrievals over the ocean using the (a) HIRAS observed spectrum and (b) the IASI observed spectrum. Note that the a priori NH3 in the title denotes the NH3 mole fraction near the surface.

    5.   Conclusions
    • In this study, a retrieval (algorithm and strategy) based on the optimal estimation method is established to derive the NH3 column from the LWIR nadir radiance spectra measured by HIRAS onboard the FY-3D satellite. We use the ASIMUT atmospheric radiative transfer model as the forward model to simulate the thermal radiation emitted from the surface and the atmosphere. The CAMS model is applied to create the a priori information of NH3. The retrieval window of 960–970 cm–1 is carefully selected to perform NH3 column retrievals from the HIRAS spectra.

      The first global map of atmospheric NH3 columns observed by the HIRAS/FY-3D satellite shows several NH3 hotspots around the world, for example, India, West Africa, and East China, where there are large NH3 emissions (Stein et al., 2014). The HIRAS NH3 columns are also compared to IASI observations. Both satellites observe similar NH3 column spatial distributions. In January 2020, the mean relative differences are within 5.5% in North America, Europe, Africa, and India, –17.2% in East China, and 35.6% in South America. The stds of their relative differences are between 46.6% and 82.4%. In July 2020, the mean relative differences are within 12% in Europe, East China, Africa, India, and 42.5% in South America. The stds of their relative differences are between 51.8% and 82.3%. These values are comparable to the differences between IASI data and collocated reference ground-based FTIR data.

      Finally, the remaining issues in the HIRAS NH3 retrievals are discussed:

      1) the SNR of the HIRAS spectrum in the high-latitude regions is too weak to perform NH3 retrieval;

      2) an improved cloud screening method is needed to better select clear-sky HIRAS pixels, especially near the ITCZ;

      3) the SNR of the HIRAS spectrum is currently too low to derive reliable NH3 columns for regions with a very low NH3 column, even in low-latitude areas such as the tropical oceans.

      Acknowledgements. The authors would like to thank Ann Carine VANDAELE (BIRA-IASB) for supporting the ASIMUT model. This study is supported by the FengYun Application Pioneering Project (FY-APP-2022.0502) and the National Natural Science Foundation of China (Grant No. 42205140).

      Data availability. The HIRAS Level 1 data is publicly available via https://www.nsmc.org.cn/. The CAMS NH3 model is publicly available via https://ads.atmosphere.copernicus.eu/cdsapp#!/dataset/cams-global-reanalysis-eac4?tab=overview. The IASI NH3 data are publicly available at https://iasi.aeris-data.fr/. The IASI L1C data are provided by the EUMETSAT https://navigator.eumetsat.int/. The HIRAS NH3 retrievals are available upon request to the authors.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return