Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-x24gv Total loading time: 0 Render date: 2024-05-21T10:10:29.284Z Has data issue: false hasContentIssue false

Part II - ‘Fluid’ Earth Applications: From the Surface to the Space

Published online by Cambridge University Press:  20 June 2023

Alik Ismail-Zadeh
Affiliation:
Karlsruhe Institute of Technology, Germany
Fabio Castelli
Affiliation:
Università degli Studi, Florence
Dylan Jones
Affiliation:
University of Toronto
Sabrina Sanchez
Affiliation:
Max Planck Institute for Solar System Research, Germany
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2023

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

References

Albers, D. J., Levine, M., Gluckman, B. et al. (2017). Personalized glucose forecasting for type 2 diabetes using data assimilation. PLoS Computational Biology, 13(4), e1005232.CrossRefGoogle ScholarPubMed
Anderson, M. G., and Burt, T. P., eds. (1985). Hydrological forecasting. Chichester: Wiley.Google Scholar
Andreadis, K. M., and Lettenmaier, D. P. (2006). Assimilating remotely sensed snow observations into a macroscale hydrology model. Advances in Water Resources, 29(6), 872–86.Google Scholar
Aoki, T., Motoyoshi, H., Kodama, Y., Yasunari, T. J., and Sugiura, K. (2007). Variations of the snow physical parameters and their effects on albedo in Sapporo, Japan. Annals of Glaciology, 46, 375–81.Google Scholar
Arsenault, K. R., Houser, P. R., De Lannoy, G. J., and Dirmeyer, P. A. (2013). Impacts of snow cover fraction data assimilation on modeled energy and moisture budgets. Journal of Geophysical Research: Atmospheres, 118(14), 7489–504.Google Scholar
Arulampalam, M. S., Maskell, S., Gordon, N., and Clapp, T. (2002). A tutorial on particle filters for online nonlinear/non-Gaussian Bayesian tracking. IEEE Transactions on Signal Processing, 50(2), 174–88.Google Scholar
Barrett, A. P. (2003). National Operational Hydrologic Remote Sensing Center SNOw Data Assimilation System (SNODAS) Products at NSIDC. NSIDC Special Report 11. Boulder, CO: National Snow and Ice Data Center. https://nsidc.org/sites/default/files/nsidc_special_report_11.pdf.Google Scholar
Bavay, M., Lehning, M., Jonas, T., and Löwe, H. (2009). Simulations of future snow cover and discharge in Alpine headwater catchments. Hydrological Processes: An International Journal, 23(1), 95108.CrossRefGoogle Scholar
Berghuijs, W. R., Woods, R. A., Hutton, C. J., and Sivapalan, M. (2016). Dominant flood generating mechanisms across the United States. Geophysical Research Letters, 43(9), 4382–90.Google Scholar
Bernier, M., Fortin, J.-P., Gauthier, Y. et al. (1999). Determination of snow water equivalent using RADARSAT SAR data in eastern Canada. Hydrological Processes, 13(18), 3041–51.Google Scholar
Bonan, G. B. (1998). The land surface climatology of the NCAR Land Surface Model coupled to the NCAR Community Climate Model. Journal of Climate, 11(6), 1307–26.2.0.CO;2>CrossRefGoogle Scholar
Brown, R. D., and Brasnett, B. (2010). Canadian Meteorological Centre (CMC) daily snow depth analysis data. Version 1 [Data Set]. Boulder, CO: National Snow and Ice Data Center. https://doi.org/10.5067/W9FOYWH0EQZ3.CrossRefGoogle Scholar
Brown, R. D., Brasnett, B., and Robinson, D. (2003). Gridded North American monthly snow depth and snow water equivalent for GCM evaluation. Atmosphere-Ocean, 41(1), 114.Google Scholar
Brun, E., David, P., Sudul, M., and Brunot, G. (1992). A numerical model to simulate snow-cover stratigraphy for operational avalanche forecasting. Journal of Glaciology, 38(128), 1322.Google Scholar
Cantet, P., Boucher, M. A., Lachance-Coutier, S., Turcotte, R., and Fortin, V. (2019). Using a particle filter to estimate the spatial distribution of the snowpack water equivalent. Journal of Hydrometeorology, 20(4), 577–94.Google Scholar
Carroll, S. S., and Carroll, T. R. (1989). Effect of uneven snow cover on airborne snow water equivalent estimates obtained by measuring terrestrial gamma radiation. Water Resources Research, 25(7), 1505–10.Google Scholar
Carroll, T. (1987). Operational airborne measurements of snow water equivalent and soil moisture using terrestrial gamma radiation in the United States. IAHS-AISH Publication, 166, 213–23.Google Scholar
Carroll, T. (2001). Airborne gamma radiation snow survey program: A user’s guide, version 5.0. National Operational Hydrologic Remote Sensing Center (NOHRSC), Chanhassen, 14.Google Scholar
Chang, A. T., Foster, J. L., and Hall, D. K. (1987). Nimbus-7 SMMR derived global snow cover parameters. Annals of Glaciology, 9, 3944.Google Scholar
Cho, E., Jacobs, J. M., and Vuyovich, C. M. (2020). The value of long‐term (40 years) airborne gamma radiation SWE record for evaluating three observation‐based gridded SWE data sets by seasonal snow and land cover classifications. Water Resources Research, 56(1). https://doi.org/10.1029/2019WR025813Google Scholar
Church, J. E. (1933). Snow surveying: Its principles and possibilities. Geographical Review, 23(4), 529–63.Google Scholar
Clark, M. P., Slater, A. G., Barrett, A. P. et al. (2006). Assimilation of snow covered area information into hydrologic and land-surface models. Advances in Water Resources, 29(8), 1209–21.CrossRefGoogle Scholar
Clifford, D. (2010). Global estimates of snow water equivalent from passive microwave instruments: History, challenges and future developments. International Journal of Remote Sensing, 31(14), 3707–26.CrossRefGoogle Scholar
Cline, D. W., Bales, R. C., and Dozier, J. (1998). Estimating the spatial distribution of snow in mountain basins using remote sensing and energy balance modeling. Water Resources Research, 34(5), 1275–85.CrossRefGoogle Scholar
Conde, V., Nico, G., Mateus, P. et al. (2019). On the estimation of temporal changes of snow water equivalent by spaceborne SAR interferometry: A new application for the Sentinel-1 mission. Journal of Hydrology and Hydromechanics, 67(1), 93100.Google Scholar
Cortés, G., Girotto, M., and Margulis, S. A. (2014). Analysis of sub-pixel snow and ice extent over the extratropical Andes using spectral unmixing of historical Landsat imagery. Remote Sensing of Environment, 141, 6478.CrossRefGoogle Scholar
Croce, P., Formichi, P., Landi, F. et al. (2018). The snow load in Europe and the climate change. Climate Risk Management, 20, 138–54.Google Scholar
De Lannoy, G. J., Reichle, R. H., Arsenault, K. R. et al. (2012). Multiscale assimilation of Advanced Microwave Scanning Radiometer–EOS snow water equivalent and Moderate Resolution Imaging Spectroradiometer snow cover fraction observations in northern Colorado. Water Resources Research, 48(1). https://doi.org/10.1029/2011WR010588.Google Scholar
De Lannoy, G. J., Reichle, R. H., Houser, P. R. et al. (2010). Satellite-scale snow water equivalent assimilation into a high-resolution land surface model. Journal of Hydrometeorology, 11(2), 352–69.Google Scholar
Dechant, C., and Moradkhani, H. (2011). Radiance data assimilation for operational snow and streamflow forecasting. Advances in Water Resources, 34(3), 351–64.Google Scholar
Deems, J. S., Painter, T. H., and Finnegan, D. C. (2013). Lidar measurement of snow depth: A review. Journal of Glaciology, 59(215), 467–79.CrossRefGoogle Scholar
Derksen, C. (2008). The contribution of AMSR-E 18.7 and 10.7 GHz measurements to improved boreal forest snow water equivalent retrievals. Remote Sensing of Environment, 112(5), 2701–10.Google Scholar
Descamps, S., Aars, J., Fuglei, E. et al. (2017). Climate change impacts on wildlife in a High Arctic archipelago – Svalbard, Norway. Global Change Biology, 23(2), 490502.Google Scholar
Deschamps-Berger, C., Gascoin, S., Berthier, E. et al. (2020). Snow depth mapping from stereo satellite imagery in mountainous terrain: Evaluation using airborne laser-scanning data. The Cryosphere, 14(9), 2925–40.Google Scholar
Dietz, A. J., Kuenzer, C., Gessner, U., and Dech, S. (2012). Remote sensing of snow: A review of available methods. International Journal of Remote Sensing, 33(13), 4094–134.Google Scholar
Domine, F., Salvatori, R., Legagneux, L. et al. (2006). Correlation between the specific surface area and the short wave infrared (SWIR) reflectance of snow. Cold Regions Science and Technology, 46(1), 60–8.Google Scholar
Dozier, J. (1989). Spectral signature of alpine snow cover from the Landsat Thematic Mapper. Remote Sensing of Environment, 28, 922.CrossRefGoogle Scholar
Durand, M., and Margulis, S. A. (2006). Feasibility test of multifrequency radiometric data assimilation to estimate snow water equivalent. Journal of Hydrometeorology, 7(3), 443–57.Google Scholar
Durand, M., and Margulis, S. A. (2007). Correcting first‐order errors in snow water equivalent estimates using a multifrequency, multiscale radiometric data assimilation scheme. Journal of Geophysical Research: Atmospheres, 112(D13). https://doi.org/10.1029/2006JD008067.Google Scholar
Durand, M., Molotch, N. P., and Margulis, S. A. (2008). A Bayesian approach to snow water equivalent reconstruction. Journal of Geophysical Research: Atmospheres, 113(D20). https://doi.org/10.1029/2008JD009894.Google Scholar
Dziubanski, D. J., and Franz, K. J. (2016). Assimilation of AMSR-E snow water equivalent data in a spatially-lumped snow model. Journal of Hydrology, 540, 2639.Google Scholar
Essery, R., Martin, E., Douville, H., Fernandez, A., and Brun, E. (1999). A comparison of four snow models using observations from an alpine site. Climate Dynamics, 15(8), 583–93.Google Scholar
Evensen, G., and Van Leeuwen, P. J. (2000). An ensemble Kalman smoother for nonlinear dynamics. Monthly Weather Review, 128(6), 1852–67.Google Scholar
Forman, B. A., and Reichle, R. H. (2014). Using a support vector machine and a land surface model to estimate large-scale passive microwave brightness temperatures over snow-covered land in North America. IEEE Journal of Selected Topics in Applied Earth Observations and Remote Sensing, 8(9), 4431–41.Google Scholar
Forman, B. A., Reichle, R. H., and Rodell, M. (2012). Assimilation of terrestrial water storage from GRACE in a snow‐dominated basin. Water Resources Research, 48(1).Google Scholar
Foster, J. L., Hall, D. K., Chang, A. T. C., and Rango, A. (1984). An overview of passive microwave snow research and results. Reviews of Geophysics, 22(2), 195208.Google Scholar
Foster, J. L., Hall, D. K., Chang, A. T. et al. (1999). Effects of snow crystal shape on the scattering of passive microwave radiation. IEEE Transactions on Geoscience and Remote Sensing, 37(2), 1165–8.CrossRefGoogle Scholar
Foster, J. L., Sun, C., Walker, J. P. et al. (2005). Quantifying the uncertainty in passive microwave snow water equivalent observations. Remote Sensing of Environment, 94(2), 187203.CrossRefGoogle Scholar
Frei, A., Tedesco, M., Lee, S. et al. (2012). A review of global satellite-derived snow products. Advances in Space Research, 50(8), 1007–29.Google Scholar
Gascoin, S., Grizonnet, M., Bouchet, M., Salgues, G., and Hagolle, O. (2019). Theia Snow collection: High-resolution operational snow cover maps from Sentinel-2 and Landsat-8 data. Earth System Science Data, 11(2), 493514.CrossRefGoogle Scholar
Girotto, M., Cortés, G., Margulis, S. A., and Durand, M. (2014a). Examining spatial and temporal variability in snow water equivalent using a 27 year reanalysis: Kern River watershed, Sierra Nevada. Water Resources Research, 50(8), 6713–34.Google Scholar
Girotto, M., Margulis, S. A., and Durand, M. (2014b). Probabilistic SWE reanalysis as a generalization of deterministic SWE reconstruction techniques. Hydrological Processes, 28(12), 3875–95.Google Scholar
Girotto, M., Musselman, K. N., and Essery, R. L. (2020). Data assimilation improves estimates of climate-sensitive seasonal snow. Current Climate Change Reports, 6, 8194.Google Scholar
Golding, D. L., and Swanson, R. H. (1978). Snow accumulation and melt in small forest openings in Alberta. Canadian Journal of Forest Research, 8(4), 380–8.Google Scholar
Günther, D., Marke, T., Essery, R., and Strasser, U. (2019). Uncertainties in snowpack simulations: Assessing the impact of model structure, parameter choice, and forcing data error on point‐scale energy balance snow model performance. Water Resources Research, 55(4), 2779–800.CrossRefGoogle Scholar
Hagopian, J., Bolcar, M., Chambers, J. et al. (2016). Advanced topographic laser altimeter system (ATLAS) receiver telescope assembly (RTA) and transmitter alignment and test. Proc. SPIE 9972, Earth Observing Systems, XXI, 997207. https://doi.org/10.1117/12.2240241.Google Scholar
Hall, D. K., Box, J. E., Casey, K. A. et al. (2008). Comparison of satellite-derived and in-situ observations of ice and snow surface temperatures over Greenland. Remote Sensing of Environment, 112(10), 3739–49.Google Scholar
Hall, D. K., and Martinec, J. (1985). Remote Sensing of Ice and Snow. London: Chapman & Hall.Google Scholar
Hall, D. K., Riggs, G. A., Salomonson, V. V., DiGirolamo, N. E., and Bayr, K. J. (2002). MODIS snow-cover products. Remote Sensing of Environment, 83(1–2), 181–94.Google Scholar
Hammond, J. C., Saavedra, F. A., and Kampf, S. K. (2018). Global snow zone maps and trends in snow persistence 2001–2016. International Journal of Climatology, 38(12), 4369–83.Google Scholar
Han, P., Long, D., Li, X. et al. (2021). A dual state-parameter updating scheme using the particle filter and high-spatial-resolution remotely sensed snow depths to improve snow simulation. Journal of Hydrology, 594, 125979.Google Scholar
Hedrick, A. R., Marks, D., Havens, S. et al. (2018). Direct insertion of NASA Airborne Snow Observatory‐derived snow depth time series into the iSnobal energy balance snow model. Water Resources Research, 54(10), 8045–63.Google Scholar
Helmert, J., Şensoy Şorman, A., Alvarado Montero, R. et al. (2018). Review of snow data assimilation methods for hydrological, land surface, meteorological and climate models: Results from a COST HarmoSnow survey. Geosciences, 8(12), 489.Google Scholar
Houser, P. R. (2013). Improved disaster management using data assimilation. In Tiefenbacher, J., ed., Approaches to Disaster Management: Examining the Implications of Hazards, Emergencies and Disasters. London: IntechOpen, 83103. https://doi.org/10.5772/3355.Google Scholar
Houser, P. R., Shuttleworth, W. J., Famiglietti, J. S. et al. (1998). Integration of soil moisture remote sensing and hydrologic modeling using data assimilation. Water Resources Research, 34(12), 3405–20.Google Scholar
Huang, C., Newman, A. J., Clark, M. P., Wood, A. W., and Zheng, X. (2017). Evaluation of snow data assimilation using the ensemble Kalman filter for seasonal streamflow prediction in the western United States. Hydrology and Earth System Sciences, 21(1), 635–50.Google Scholar
Hüsler, F., Jonas, T., Riffler, M., Musial, J. P., and Wunderle, S. (2014). A satellite-based snow cover climatology (1985–2011) for the European Alps derived from AVHRR data. The Cryosphere, 8(1), 7390.Google Scholar
Jepsen, S. M., Harmon, T. C., Ficklin, D. L., Molotch, N. P., and Guan, B. (2018). Evapotranspiration sensitivity to air temperature across a snow-influenced watershed: Space-for-time substitution versus integrated watershed modeling. Journal of Hydrology, 556, 645–59.Google Scholar
Jordan, R. E. (1991). A One-Dimensional Temperature Model for a Snow Cover: Technical Documentation for SNTHERM. 89. (No. CRREL-SR-91-16). Hanover, NH: Cold Regions Research and Engineering Lab,Google Scholar
Jost, G., Weiler, M., Gluns, D. R., and Alila, Y. (2007). The influence of forest and topography on snow accumulation and melt at the watershed-scale. Journal of Hydrology, 347(1–2), 101–15.Google Scholar
Kelly, R. (2009). The AMSR-E snow depth algorithm: Description and initial results. Journal of the Remote Sensing Society of Japan, 29(1), 307–17.Google Scholar
Kendra, J. R., Sarabandi, K., and Ulaby, F. T. (1998). Radar measurements of snow: Experiment and analysis. IEEE Transactions on Geoscience and Remote Sensing, 36(3), 864–79.Google Scholar
Kim, E., Gatebe, C., Hall, D., et al. (2017). NASA’s SnowEx campaign: Observing seasonal snow in a forested environment. 2017 IEEE International Geoscience and Remote Sensing Symposium (IGARSS), Fort Worth, TX, USA, July 23–28, 2017, pp. 1388–90, https://doi.org/10.1109/IGARSS.2017.8127222.Google Scholar
Kinar, N. J., and Pomeroy, J. W. (2015). Measurement of the physical properties of the snowpack. Reviews of Geophysics, 53(2), 481544.Google Scholar
Krinner, G., Derksen, C., Essery, R. et al. (2018). ESM-SnowMIP: Assessing snow models and quantifying snow-related climate feedbacks. Geoscientific Model Development, 11(12), 5027–49.Google Scholar
Kumar, M., Marks, D., Dozier, J., Reba, M., and Winstral, A. (2013). Evaluation of distributed hydrologic impacts of temperature-index and energy-based snow models. Advances in Water Resources, 56, 7789.Google Scholar
Lafaysse, M., Cluzet, B., Dumont, M. et al. (2017). A multiphysical ensemble system of numerical snow modelling. The Cryosphere, 11(3), 1173–98.Google Scholar
Largeron, C., Dumont, M., Morin, S. et al. (2020). Toward snow cover estimation in mountainous areas using modern data assimilation methods: A review. Frontiers in Earth Science, 8, 325. https://doi.org/10.3389/feart.2020.00325.Google Scholar
Lehning, M., Bartelt, P., Brown, B. et al. (1999). SNOWPACK model calculations for avalanche warning based upon a new network of weather and snow stations. Cold Regions Science and Technology, 30(1–3), 145–57.Google Scholar
Leisenring, M., and Moradkhani, H. (2011). Snow water equivalent prediction using Bayesian data assimilation methods. Stochastic Environmental Research and Risk Assessment, 25(2), 253–70.Google Scholar
Lemmetyinen, J., Derksen, C., Rott, H. et al. (2018). Retrieval of effective correlation length and snow water equivalent from radar and passive microwave measurements. Remote Sensing, 10(2), 170. https://doi.org/10.3390/rs10020170.Google Scholar
Li, D., Durand, M., and Margulis, S. A. (2012). Potential for hydrologic characterization of deep mountain snowpack via passive microwave remote sensing in the Kern River basin, Sierra Nevada, USA. Remote Sensing of Environment, 125, 3448.Google Scholar
Li, D., Lettenmaier, D. P., Margulis, S. A., and Andreadis, K. (2019). The value of accurate high-resolution and spatially continuous snow information to streamflow forecasts. Journal of Hydrometeorology, 20(4), 731–49.Google Scholar
Lievens, H., Demuzere, M., Marshall, H.-P. et al. (2019). Snow depth variability in the Northern Hemisphere mountains observed from space. Nature Communications, 10(1), 112.Google Scholar
Liu, Y., Fang, Y., and Margulis, S. A. (2021). Spatiotemporal distribution of seasonal snow water equivalent in High Mountain Asia from an 18-year Landsat-MODIS era snow reanalysis dataset. The Cryosphere, 15, 5261–80.Google Scholar
Liu, Y., Peters-Lidard, C. D., Kumar, S. et al. (2013). Assimilating satellite-based snow depth and snow cover products for improving snow predictions in Alaska. Advances in Water Resources, 54, 208–27.Google Scholar
Lundquist, J. D., Chickadel, C., Cristea, N. et al. (2018). Separating snow and forest temperatures with thermal infrared remote sensing. Remote Sensing of Environment, 209, 764–79.Google Scholar
Lundquist, J., Hughes, M., Gutmann, E., and Kapnick, S. (2019). Our skill in modeling mountain rain and snow is bypassing the skill of our observational networks. Bulletin of the American Meteorological Society, 100(12), 2473–90.Google Scholar
Luojus, K., Pulliainen, J., Takala, M. et al. (2021). GlobSnow v3. 0 Northern Hemisphere snow water equivalent dataset. Scientific Data, 8(1), 116.Google Scholar
Lv, Z., and Pomeroy, J. W. (2020). Assimilating snow observations to snow interception process simulations. Hydrological Processes, 34(10), 2229–46.Google Scholar
Magnusson, J., Gustafsson, D., Hüsler, F., and Jonas, T. (2014). Assimilation of point SWE data into a distributed snow cover model comparing two contrasting methods. Water Resources Research, 50(10), 7816–35.Google Scholar
Margulis, S. A., Cortés, G., Girotto, M., and Durand, M. (2016). A Landsat-era Sierra Nevada snow reanalysis (1985–2015). Journal of Hydrometeorology, 17(4), 1203–21.Google Scholar
Margulis, S. A., Fang, Y., Li, D., Lettenmaier, D. P., and Andreadis, K. (2019). The utility of infrequent snow depth images for deriving continuous space‐time estimates of seasonal snow water equivalent. Geophysical Research Letters, 46(10), 5331–40.Google Scholar
Margulis, S. A., Girotto, M., Cortés, G., and Durand, M. (2015). A particle batch smoother approach to snow water equivalent estimation. Journal of Hydrometeorology, 16(4), 1752–72.Google Scholar
Meromy, L., Molotch, N. P., Link, T. E., Fassnacht, S. R., and Rice, R. (2013). Subgrid variability of snow water equivalent at operational snow stations in the western USA. Hydrological Processes, 27(17), 2383–400.Google Scholar
Miller, S. D., Lee, T. F., and Fennimore, R. L. (2005). Satellite-based imagery techniques for daytime cloud/snow delineation from MODIS. Journal of Applied Meteorology, 44(7), 987–97.Google Scholar
Molotch, N. P., and Margulis, S. A. (2008). Estimating the distribution of snow water equivalent using remotely sensed snow cover data and a spatially distributed snowmelt model: A multi-resolution, multi-sensor comparison. Advances in Water Resources, 31(11), 1503–14.Google Scholar
Musselman, K. N., Addor, N., Vano, J. A., and Molotch, N. P. (2021). Winter melt trends portend widespread declines in snow water resources. Nature Climate Change, 11(5), 418–24.Google Scholar
Musselman, K. N., Lehner, F., Ikeda, K. et al. (2018). Projected increases and shifts in rain-on-snow flood risk over western North America. Nature Climate Change, 8(9), 808–12.Google Scholar
Musselman, K. N., Pomeroy, J. W., Essery, R. L., and Leroux, N. (2015). Impact of windflow calculations on simulations of alpine snow accumulation, redistribution and ablation. Hydrological Processes, 29(18), 3983–99.Google Scholar
Nadim, F., Kjekstad, O., Peduzzi, P., Herold, C., and Jaedicke, C. (2006). Global landslide and avalanche hotspots. Landslides, 3(2), 159–73.Google Scholar
Nagler, T., Rott, H., Ripper, E., Bippus, G., and Hetzenecker, M. (2016). Advancements for snowmelt monitoring by means of Sentinel-1 SAR. Remote Sensing, 8(4), 348. https://doi.org/10.3390/rs8040348.Google Scholar
Niu, G.-Y., Yang, Z.-L., Mitchell, K. E. et al. (2011). The community Noah land surface model with multiparameterization options (Noah‐MP): 1. Model description and evaluation with local‐scale measurements. Journal of Geophysical Research: Atmospheres, 116(D12).Google Scholar
Nolan, M., Larsen, C., and Sturm, M. (2015). Mapping snow depth from manned aircraft on landscape scales at centimeter resolution using structure-from-motion photogrammetry. The Cryosphere, 9(4), 1445–63.Google Scholar
Oaida, C. M., Reager, J. T., Andreadis, K. M. et al. (2019). A high-resolution data assimilation framework for snow water equivalent estimation across the western United States and validation with the airborne snow observatory. Journal of Hydrometeorology, 20(3), 357–78.Google Scholar
Ohmura, A. (2001). Physical basis for the temperature-based melt-index method. Journal of Applied Meteorology, 40(4), 753–61.Google Scholar
Painter, T. H., Berisford, D. F., Boardman, J. W. et al. (2016). The Airborne Snow Observatory: Fusion of scanning lidar, imaging spectrometer, and physically-based modeling for mapping snow water equivalent and snow albedo. Remote Sensing of Environment, 184, 139–52.Google Scholar
Painter, T. H., Bryant, A. C., and Skiles, S. M. (2012). Radiative forcing by light absorbing impurities in snow from MODIS surface reflectance data. Geophysical Research Letters, 39(17).Google Scholar
Painter, T. H., Rittger, K., McKenzie, C. et al. (2009). Retrieval of subpixel snow covered area, grain size, and albedo from MODIS. Remote Sensing of Environment, 113(4), 868–79.Google Scholar
Picard, G., Sandells, M., and Löwe, H. (2018). SMRT: An active–passive microwave radiative transfer model for snow with multiple microstructure and scattering formulations (v1. 0). Geoscientific Model Development, 11(7), 2763–88.Google Scholar
Pulliainen, J., Luojus, K., Derksen, C. et al. (2020). Patterns and trends of Northern Hemisphere snow mass from 1980 to 2018. Nature, 581(7808), 294–8.Google Scholar
Raleigh, M. S., Livneh, B., Lapo, K., and Lundquist, J. D. (2016). How does availability of meteorological forcing data impact physically based snowpack simulations? Journal of Hydrometeorology, 17(1), 99120.Google Scholar
Rango, A. (1996). Spaceborne remote sensing for snow hydrology applications. Hydrological Sciences Journal, 41(4), 477–94.Google Scholar
Rice, R., Bales, R. C., Painter, T. H., and Dozier, J. (2011). Snow water equivalent along elevation gradients in the Merced and Tuolumne River basins of the Sierra Nevada. Water Resources Research, 47(8).Google Scholar
Riggs, G. A., Hall, D. K., and Román, M. O. (2017). Overview of NASA’s MODIS and visible infrared imaging radiometer suite (VIIRS) snow-cover earth system data records. Earth System Science Data, 9(2), 765–77.CrossRefGoogle Scholar
Rodell, M., and Houser, P. R. (2004). Updating a land surface model with MODIS-derived snow cover. Journal of Hydrometeorology, 5(6), 1064–75.Google Scholar
Rosenthal, W., and Dozier, J. (1996). Automated mapping of montane snow cover at subpixel resolution from the Landsat Thematic Mapper. Water Resources Research, 32(1), 115–30.Google Scholar
Santi, E., Brogioni, M., Leduc-Leballeur, M. et al. (2022). Exploiting the ANN Potential in Estimating Snow Depth and Snow Water Equivalent From the Airborne SnowSAR Data at X-and Ku-Bands. IEEE Transactions on Geoscience and Remote Sensing, 60, 116. https://do.org/ 10.1109/TGRS.2021.3086893.Google Scholar
Schlosser, C. A., Robock, A., Vinnikov, K. Y., Speranskaya, N. A., and Xue, Y. (1997). 18-year land-surface hydrology model simulations for a midlatitude grassland catchment in Valdai, Russia. Monthly Weather Review, 125(12), 3279–96.Google Scholar
Schmidt, R. A. (1982). Properties of blowing snow. Reviews of Geophysics, 20(1), 3944.Google Scholar
Schmugge, T. J., Kustas, W. P., Ritchie, J. C., Jackson, T. J., and Rango, A. (2002). Remote sensing in hydrology. Advances in Water Resources, 25(8–12), 1367–85.Google Scholar
Serreze, M. C., Clark, M. P., Armstrong, R. L., McGinnis, D. A., and Pulwarty, R. S. (1999). Characteristics of the western United States snowpack from snowpack telemetry (SNOTEL) data. Water Resources Research, 35(7), 2145–60.Google Scholar
Shi, J., and Dozier, J. (2000). Estimation of snow water equivalence using SIR-C/X-SAR. I. Inferring snow density and subsurface properties. IEEE Transactions on Geoscience and Remote Sensing, 38(6), 2465–74.Google Scholar
Shufen, S., and Yongkang, X. (2001). Implementing a new snow scheme in simplified simple biosphere model. Advances in Atmospheric Sciences, 18(3), 335–54.Google Scholar
Skiles, S. M., Flanner, M., Cook, J. M., Dumont, M., and Painter, T. H. (2018). Radiative forcing by light-absorbing particles in snow. Nature Climate Change, 8(11), 964–71.Google Scholar
Skiles, S. M., and Painter, T. H. (2019). Toward understanding direct absorption and grain size feedbacks by dust radiative forcing in snow with coupled snow physical and radiative transfer modeling. Water Resources Research, 55(8), 7362–78.Google Scholar
Slater, A. G., Barrett, A. P., Clark, M. P., Lundquist, J. D., and Raleigh, M. S. (2013). Uncertainty in seasonal snow reconstruction: Relative impacts of model forcing and image availability. Advances in Water Resources, 55, 165–77.Google Scholar
Slater, A. G., and Clark, M. P. (2006). Snow data assimilation via an ensemble Kalman filter. Journal of Hydrometeorology, 7(3), 478–93.Google Scholar
Smith, C. D., Kontu, A., Laffin, R., and Pomeroy, J. W. (2017). An assessment of two automated snow water equivalent instruments during the WMO Solid Precipitation Intercomparison Experiment. The Cryosphere, 11(1), 101–16.Google Scholar
Smyth, E. J., Raleigh, M. S., and Small, E. E. (2019). Particle filter data assimilation of monthly snow depth observations improves estimation of snow density and SWE. Water Resources Research, 55(2), 1296–311.Google Scholar
Stigter, E. E., Wanders, N., Saloranta, T. M. et al. (2017). Assimilation of snow cover and snow depth into a snow model to estimate snow water equivalent and snowmelt runoff in a Himalayan catchment. The Cryosphere, 11(4), 1647–64.Google Scholar
Strozzi, T., and Matzler, C. (1998). Backscattering measurements of alpine snowcovers at 5.3 and 35 GHz. IEEE Transactions on Geoscience and Remote Sensing, 36(3), 838–48.Google Scholar
Su, H., Yang, Z.-L., Niu, G.-Y., and Dickinson, R. E. (2008). Enhancing the estimation of continental‐scale snow water equivalent by assimilating MODIS snow cover with the ensemble Kalman filter. Journal of Geophysical Research: Atmospheres, 113(D8).Google Scholar
Tachiiri, K., Shinoda, M., Klinkenberg, B., and Morinaga, Y. (2008). Assessing Mongolian snow disaster risk using livestock and satellite data. Journal of Arid Environments, 72(12), 2251–63.Google Scholar
Tapley, B. D., Bettadpur, S., Ries, J. C., Thompson, P. F., and Watkins, M. M. (2004). GRACE measurements of mass variability in the Earth system. Science, 305(5683), 503–5.Google Scholar
Tedesco, M. (2014). Remote Sensing of the Cryosphere. Chichester: John Wiley & Sons.Google Scholar
Tedesco, M., Pulliainen, J., Takala, M., Hallikainen, M., and Pampaloni, P. (2004). Artificial neural network-based techniques for the retrieval of SWE and snow depth from SSM/I data. Remote Sensing of Environment, 90(1), 7685.Google Scholar
Thirel, G., Salamon, P., Burek, P., and Kalas, M. (2013). Assimilation of MODIS snow cover area data in a distributed hydrological model using the particle filter. Remote Sensing, 5(11), 5825–50.Google Scholar
Todt, M., Rutter, N., Fletcher, C. G. et al. (2018). Simulation of longwave enhancement in boreal and montane forests. Journal of Geophysical Research: Atmospheres, 123(24), 13731–47.Google Scholar
Toure, A. M., Reichle, R. H., Forman, B. A., Getirana, A., and De Lannoy, G. J. (2018). Assimilation of MODIS snow cover fraction observations into the NASA catchment land surface model. Remote Sensing, 10(2), 316.Google Scholar
Van Leeuwen, P. J. (2009). Particle filtering in geophysical systems. Monthly Weather Review, 137(12), 4089–114.Google Scholar
Veyssière, G., Karbou, F., Morin, S., Lafaysse, M., and Vionnet, V. (2019). Evaluation of sub-kilometric numerical simulations of c-band radar backscatter over the French alps against sentinel-1 observations. Remote Sensing, 11(1), 8. https://doi.org/10.3390/rs11010008.Google Scholar
Viallon-Galinier, L., Hagenmuller, P., and Lafaysse, M. (2020). Forcing and evaluating detailed snow cover models with stratigraphy observations. Cold Regions Science and Technology, 180, 103163.Google Scholar
Viviroli, D., Dürr, H. H., Messerli, B., Meybeck, M., and Weingartner, R. (2007). Mountains of the world, water towers for humanity: Typology, mapping, and global significance. Water Resources Research, 43(7).Google Scholar
Wang, J., Forman, B. A., Girotto, M., and Reichle, R. H. (2021). Estimating terrestrial snow mass via multi‐sensor assimilation of synthetic AMSR‐E brightness temperature spectral differences and synthetic GRACE terrestrial water storage retrievals. Water Resources Research, 57(9), e2021WR029880.Google Scholar
Wiesmann, A., and Mätzler, C. (1999). Microwave emission model of layered snowpacks. Remote Sensing of Environment, 70(3), 307316.Google Scholar
Winstral, A., Magnusson, J., Schirmer, M., and Jonas, T. (2019). The bias‐detecting ensemble: A new and efficient technique for dynamically incorporating observations into physics‐based, multilayer snow models. Water Resources Research, 55(1), 613–31.Google Scholar
Xue, Y., and Forman, B. A. (2017). Integration of satellite-based passive microwave brightness temperature observations and an ensemble-based land data assimilation framework to improve snow estimation in forested regions. 2017 IEEE International Geoscience and Remote Sensing Symposium (IGARSS), 311–14.Google Scholar

References

Arthern, R. J. (2015). Exploring the use of transformation group priors and the method of maximum relative entropy for Bayesian glaciological inversions. Journal of Glaciology, 61(229), 947–62.Google Scholar
Arthern, R. J., and Gudmundsson, G. H. (2010). Initialization of ice-sheet forecasts viewed as an inverse Robin Problem, Journal of Glaciology, 56(197), 527–33.Google Scholar
Ashmore, D. W., Bingham, R. G., Ross, N. et al. (2020). Englacial architecture and age-depth constraints across the West Antarctic ice sheet. Geophysical Research Letters, 47(6), e2019GL086663. https://doi.org/10.1029/2019GL086663.Google Scholar
Babaniyi, O., Nicholson, R., Villa, U., and Petra, N. (2021). Inferring the basal sliding coefficient field for the stokes ice sheet model under rheological uncertainty. The Cryosphere, 15(4), 1731–50.Google Scholar
Barnes, J. M., dos Santos, T. D., Goldberg, D., Gudmundsson, G. H., Morlighem, M., and De Rydt, J. (2021). The transferability of adjoint inversion products between different ice flow models. The Cryosphere, 15(4), 19752000. https://doi.org/10.5194/tc-15-1975-2021.Google Scholar
Bell, R. E., et al. (2011). Widespread persistent thickening of the east Antarctic Ice Sheet by freezing from the base. Science, 331(6024), 1592–5. https://doi.org/10.1126/science.1200109.Google Scholar
Blatter, H. (1995). Velocity and stress-fields in grounded glaciers: A simple algorithm for including deviatoric stress gradients. Journal of Glaciology, 41(138), 333–44.Google Scholar
Bodart, J. A., Bingham, R. G., Ashmore, D. et al. (2021). Age-depth stratigraphy of Pine Island Glacier inferred from airborne radar and ice-core chronology. Journal of Geophysical Research: Earth Surface, 126(4), e2020JF005,927.Google Scholar
Borstad, C. P., Rignot, E., Mouginot, J., and Schodlok, M. P. (2013). Creep deformation and buttressing capacity of damaged ice shelves: Theory and application to Larsen C ice shelf. The Cryosphere, 7, 1931–47. https://doi.org/10.5194/tc-7-1931-2013.Google Scholar
Brinkerhoff, D., Aschwanden, A., and Fahnestock, M. (2021). Constraining subglacial processes from surface velocity observations using surrogate-based Bayesian inference. Journal of Glaciology, 67(263), 385403.Google Scholar
Bryson, A. E., and Ho, Y.-C. (2018). Applied Optimal Control: Optimization, Estimation, and Control. Boca Raton, FL:CRC Press.Google Scholar
Budd, W. F., and Warner, R. C. (1996). A computer scheme for rapid calculations of balance-flux distributions. Annals of Glaciology, 23, 2127.Google Scholar
Bui-Thanh, T., Ghattas, O., Martin, J., and Stadler, G.. (2013). A computational framework for infinite-dimensional Bayesian inverse problems Part I: The linearized case, with application to global seismic inversion. SIAM Journal on Scientific Computing, 35(6). https://doi.org/10.1137/12089586X.Google Scholar
Chandler, D. M., Hubbard, A. L., Hubbard, B., and Nienow, P. (2006). A Monte Carlo error analysis for basal sliding velocity calculations. Journal of Geophysical Research: Earth Surface, 111(F4). https://doi.org/10.1029/2006JF000476.Google Scholar
Chatfield, C. (1995). Model uncertainty, data mining and statistical inference. Journal of the Royal Statistical Society: Series A (Statistics in Society), 158(3), 419–44.Google Scholar
Clarke, G. K. C., Berthier, E., Schoof, C. G., and Jarosch, A. H. (2009). Neural networks applied to estimating subglacial topography and glacier volume. Journal of Climate, 22(8), 2146–60. https://doi.org/10.1175/2008JCLI2572.1.Google Scholar
Conway, H., Catania, G., Raymond, C. F., Gades, A. M., Scambos, T. A., and Engelhardt, H. (2002). Switch of flow direction in an Antarctic ice stream. Nature, 419(6906), 465–67. https://doi.org/10.1038/nature01081.Google Scholar
Cornford, S. L., et al. (2015). Century-scale simulations of the response of the West Antarctic Ice Sheet to a warming climate. The Cryosphere, 9, 1579–600. https://doi.org/10.5194/tc-9-1579-2015.Google Scholar
Cuffey, K. M., and Paterson, W. S. B. (2010). The Physics of Glaciers, 4th ed. Oxford: Elsevier.Google Scholar
Dansgaard, W., Johnsen, S. J., Møller, J., and Langway, C. C. (1969). One thousand centuries of climatic record from camp century on the Greenland ice sheet. Science, 166(3903), 377–80.Google Scholar
Edwards, T., et al. (2021). Quantifying uncertainties in the land ice contribution to sea level rise this century. Nature, 593. https://doi.org/10.1038/s41586-021-03302-y.Google Scholar
Evans, S., and Robin, G. d. Q. (1966). Glacier depth-sounding from air. Nature, 210(5039), 883–5. https://doi.org/10.1038/210883a0.Google Scholar
Farinotti, D., Huss, M., Bauder, A., Funk, M., and Truffer, M. (2009). A method to estimate the ice volume and ice-thickness distribution of alpine glaciers. Journal of Glaciology, 55(191), 422–30.Google Scholar
Farinotti, D., et al. (2017). How accurate are estimates of glacier ice thickness? Results from ITMIX, the Ice Thickness Models Intercomparison eXperiment. The Cryosphere, 11(2), 949–70. https://doi.org/10. 5194/tc-11–949–2017.Google Scholar
Farinotti, D., et al. (2021). Results from the Ice Thickness Models Intercomparison eXperiment Phase 2 (ITMIX2). Frontiers in Earth Science, 8, 484. https://doi.org/10.3389/feart.2020.571923.Google Scholar
Fastook, J. L., Brecher, H. H., and Hughes, T. J. (1995). Derived bedrock elevations, strain rates and stresses from measured surface elevations and velocities: Jakobshavns Isbrae, Greenland. Journal of Glaciology, 41(137), 161–73.Google Scholar
Furst, J. J., Durand, G., Gillet-Chaulet, F. et al. (2015). Assimilation of Antarctic velocity observations provides evidence for uncharted pinning points. The Cryosphere, 9, 1427–43. https://doi.org/10.5194/tc-9-1427-2015.Google Scholar
Fürst, J. J., et al. (2017). Application of a two-step approach for mapping ice thickness to various glacier types on Svalbard. The Cryosphere, 11(5), 2003–32. https://doi.org/10.5194/tc-11-2003-2017.Google Scholar
Fürst, J. J., et al. (2018). The ice-free topography of Svalbard. Geophysical Research Letters, 45(21), 11,7609. https://doi.org/10.1029/2018GL079734.Google Scholar
Giering, R., Kaminski, T., and Slawig, T. (2005). Generating efficient derivative code with TAF. Future Generation Computer Systems, 21(8), 1345–55. https://doi.org/10.1016/j.future.2004.11.003.Google Scholar
Gilbert, J. C., and Lemaréchal, C. (1989). Some numerical experiments with variable-storage quasi-Newton algorithms. Mathematical Programming, 45(1–3), 407–35. https://doi.org/10.1007/BF01589113.Google Scholar
Gillet-Chaulet, F. (2020). Assimilation of surface observations in a transient marine ice sheet model using an ensemble Kalman filter. The Cryosphere, 14, 811–32. https://doi.org/10.5194/tc-14-811-2020.Google Scholar
Gillet-Chaulet, F., Gagliardini, O., Seddik, H. et al. (2012). Greenland Ice Sheet contribution to sea-level rise from a new-generation ice-sheet model. The Cryosphere, 6, 1561–76. https://doi.org/10.5194/tc-6-1561-2012.Google Scholar
Glen, J. W. (1955). The creep of polycrystalline ice. Proceedings of the Royal Society A, 228(1175), 519–38.Google Scholar
Goldberg, D. N., and Heimbach, P. (2013). Parameter and state estimation with a time-dependent adjoint marine ice sheet model. The Cryosphere, 17, 1659–78.Google Scholar
Goldberg, D. N., Heimbach, P., Joughin, I., and Smith, B. (2015). Committed retreat of Smith, Pope, and Kohler Glaciers over the next 30 years inferred by transient model calibration. The Cryosphere, 9(6), 2429–46. https://doi.org/10.5194/tc-9-2429-2015.Google Scholar
Goldberg, D. N., Narayanan, S. H. K., Hascoet, L., and Utke, J. (2016). An optimized treatment for algorithmic differentiation of an important glaciological fixed-point problem. Geoscientific Model Development, 9(5), 1891–904.Google Scholar
Greve, R., and Blatter, H. (2009). Dynamics of Ice Sheets and Glaciers. Berlin: Springer Science & Business Media.Google Scholar
Greve, R., Saito, F., and Abe-Ouchi, A. (2011). Initial results of the SeaRISE numerical experiments with the models SICOPOLIS and IcIES for the Greenland ice sheet. Annals of Glaciology, 52(58), 2330.Google Scholar
Griewank, A., and Walther, A. (2008). Evaluating Derivatives: Principles and Techniques of Algorithmic Differentiation, vol. 19, 2nd ed. Philadelphia, PA: SIAM Frontiers in Applied Mathematics.Google Scholar
Gudmundsson, G. H. (2003). Transmission of basal variability to a glacier surface. Journal of Geophysical Research: Solid Earth, 108(B5), 119. https://doi.org/10.1029/2002JB002107.Google Scholar
Habermann, M., Truffer, M., and Maxwell, D. (2013). Changing basal conditions during the speed-up of Jakobshavn Isbræ, Greenland. The Cryosphere, 7(6), 1679–92. https://doi.org/10.5194/tc-7-1679-2013.Google Scholar
Hadamard, J. (1902). Sur les probl`emes aux Dérivées partielles et leur signification physique, Princeton University Bulletin, 13, 4952.Google Scholar
Hansen, P. C. (2000). The L-curve and its use in the numerical treatment of inverse problems. In Johnston, P., ed., Computational Inverse Problems in Electrocardiology: Advances in Computational Bioengineering. Southampton: WIT Press, pp. 119–42.Google Scholar
Heimbach, P., and Bugnion, V. (2009). Greenland ice-sheet volume sensitivity to basal, surface and initial conditions derived from an adjoint model. Annals of Glaciology, 50(52), 6780.Google Scholar
Hindmarsh, R. C., Leysinger-Vieli, G. J.-M., and Parrenin, F. (2009). A large-scale numerical model for computing isochrone geometry. Annals of Glaciology, 50(51), 130–40.Google Scholar
Huss, M., and Farinotti, D. (2012). Distributed ice thickness and volume of all glaciers around the globe. Journal of Geophysical Research: Earth Surface, 117(F4). https://doi.org/10.1029/2012JF002523.Google Scholar
Hutter, K. (1983). Theoretical Glaciology: Material Science of Ice and the Mechanics of Glaciers and Ice Sheets. Dordrecht: D. Reidel Publishing.Google Scholar
Isaac, T., Petra, N., Stadler, G., and Ghattas, O. (2015). Scalable and efficient algorithms for the propagation of uncertainty from data through inference to prediction for large-scale problems, with application to flow of the Antarctic ice sheet. Journal of Computational Physics, 296, 348–38.Google Scholar
Joughin, I., MacAyeal, D. R., and Tulaczyk, S. (2004). Basal shear stress of the Ross ice streams from control method inversions. Journal of Geophysical Research: Solid Earth, 109(B9), 162. https://doi.org/10.1029/2003JB002960.Google Scholar
Joughin, I., Smith, B. E., and Howat, I. M. (2018). A complete map of Greenland ice velocity derived from satellite data collected over 20 years. Journal of Glaciology, 64(243), 111. https://doi.org/10.1017/jog.2017.73.Google Scholar
Jouvet, G. (2023). Inversion of a Stokes glacier flow model emulated by deep learning. Journal of Glaciology, 69(273), 1326.Google Scholar
Jouvet, G., Cordonnier, G., Kim, B., Lüthi, M., Vieli, A. and Aschwanden, A. (2022). Deep learning speeds up ice flow modelling by several orders of magnitude. Journal of Glaciology, 68(270), 65164.Google Scholar
Kalmikov, A. G., and Heimbach, P. (2014). A hessian-based method for uncertainty quantification in global ocean state estimation. SIAM Journal on Scientific Computing, 36(5), S267S295.Google Scholar
Karlsson, N. B., Bingham, R. G., Rippin, D. M. et al. (2014). Constraining past accumulation in the central Pine Island Glacier basin, West Antarctica, using radio-echo sounding. Journal of Glaciology, 60(221), 553–62.Google Scholar
Kohn, R., and Vogelius, M. (1984). Determining conductivity by boundary measurements. Communications on Pure and Applied Mathematics, 37(3), 289–98. https://doi.org/10.1002/cpa. 3160370302.Google Scholar
Koutnik, M. R., and Waddington, E. D. (2012). Well-posed boundary conditions for limited-domain models of transient ice flow near an ice divide. Journal of Glaciology, 58(211), 1008–20.Google Scholar
Koutnik, M. R., Fudge, T., Conway, H. et al. (2016). Holocene accumulation and ice flow near the West Antarctic ice sheet divide ice core site. Journal of Geophysical Research: Earth Surface, 121(5), 907924.Google Scholar
Koziol, C. P., Todd, J. A., Goldberg, D. N., and Maddison, J. R. (2021). fenics ice 1.0: A framework for quantifying initialization uncertainty for time-dependent ice sheet models. Geoscientific Model Development, 14, 5843–61. https://doi.org/10.5194/gmd-14-5843-2021.Google Scholar
Krasnopolsky, V. M., and Schiller, H. (2003). Some neural network applications in environmental sciences. Part I: Forward and inverse problems in geophysical remote measurements. Neural Networks, 16(34), 321–34.Google Scholar
Kyrke-Smith, T. M., Gudmundsson, G. H., and Farrell, P. E. (2017). Can seismic observations of bed conditions on ice streams help constrain parameters in ice flow models? Journal of Geophysical Research: Earth Surface, 122(11), 2269–82.Google Scholar
Larour, E. (2005). Modélisation numérique du comportement des banquises flottantes, validée par imagerie satellitaire, Ph.D. thesis, Ecole Centrale Paris.Google Scholar
Larour, E., Utke, J., Csatho, B. et al. (2014). Inferred basal friction and surface mass balance of the Northeast Greenland Ice Stream using data assimilation of ICESat (Ice Cloud and land Elevation Satellite) surface altimetry and ISSM (Ice Sheet System Model). The Cryosphere, 8(6), 2335–51. https://doi.org/10.5194/tc-8–2335–2014.Google Scholar
Lee, V., Cornford, S. L., and Payne, A. J. (2015). Initialization of an ice-sheet model for present-day Greenland. Annals of Glaciology, 56(70), 129–40. https://doi.org/10.3189/2015AoG70A121.Google Scholar
Leong, W. J., and Horgan, H. J. (2020). DeepBedMap: A deep neural network for resolving the bed topography of Antarctica. The Cryosphere, 14(11), 3687–705. https://doi.org/10.5194/tc-14-3687-2020.Google Scholar
Li, D., Gurnis, M., and Stadler, G. (2017). Towards adjoint-based inversion of time-dependent mantle convection with nonlinear viscosity. Geophysical Journal International, 209(1), 86105. https://doi.org/10.1093/gji/ggw493.Google Scholar
Logan, L. C., Narayanan, S. H. K., Greve, R., and Heimbach, P. (2020). Sicopolis-ad v1: an open-source adjoint modeling framework for ice sheet simulation enabled by the algorithmic differentiation tool OpenAD. Geoscientific Model Development, 13(4), 1845–64.Google Scholar
Lorenc, A. C. (2003). The potential of the ensemble Kalman filter for NWP: A comparison with 4d-var. Quarterly Journal of the Royal Meteorological Society: A Journal of the Atmospheric Sciences, Applied Meteorology and Physical Oceanography, 129(595), 3183–203.Google Scholar
MacAyeal, D. R. (1989). Large-scale ice flow over a viscous basal sediment: Theory and application to Ice Stream B, Antarctica. Journal of Geophysical Research Atmospheres, 94(B4), 4071–87.Google Scholar
MacAyeal, D. R. (1992a). Irregular oscillations of the West Antarctic ice sheet. Nature, 359(6390), 2932.Google Scholar
MacAyeal, D. R. (1992b). The basal stress distribution of Ice Stream E, Antarctica, inferred by control methods. Journal of Geophysical Research: Solid Earth, 97(B1), 595603.Google Scholar
MacAyeal, D. R. (1993). A tutorial on the use of control methods in ice-sheet modelling. Journal of Glaciology, 39(131), 91–8.Google Scholar
MacGregor, J. A. et al. (2015). Radiostratigraphy and age structure of the Greenland Ice Sheet. Journal of Geophysical Research: Earth Surface, 120(2), 212–41. https://doi.org/10.1002/2014JF003215.Google Scholar
Maddison, J. R., Goldberg, D. N., and Goddard, B. D. (2019). Automated calculation of higher order partial differential equation constrained derivative information. SIAM Journal on Scientific Computing, 41(5), C417C445.Google Scholar
Martin, J., Wilcox, L. C., Burstedde, C., and Ghattas, O. (2012). A stochastic Newton MCMC method for large-scale statistical inverse problems with application to seismic inversion. SIAM Journal on Scientific Computing, 34(3), A1460A1487.Google Scholar
Michel, L., Picasso, M., Farinotti, D., Funk, M., and Blatter, H. (2014). Estimating the ice thickness of shallow glaciers from surface topography and mass-balance data with a shape optimization algorithm. Computers & Geosciences, 66, 182–99.Google Scholar
Morlighem, M., Rignot, E., Seroussi, H. et al. (2010). Spatial patterns of basal drag inferred using control methods from a full-Stokes and simpler models for Pine Island Glacier, West Antarctica. Geophysical Research Letters, 37(L14502), 16. https://doi.org/10.1029/2010GL043853.Google Scholar
Morlighem, M., Rignot, E., Seroussi, H. et al. (2011). A mass conservation approach for mapping glacier ice thickness. Geophysical Research Letters, 38(L19503), 16. https://doi.org/10.1029/2011GL048659.Google Scholar
Morlighem, M., Rignot, E., Mouginot, J. et al. (2013a). High-resolution bed topography mapping of Russell Glacier, Greenland, inferred from Operation IceBridge data. Journal of Glaciology, 59(218), 1015–23. https://doi.org/10.3189/2013JoG12J235.Google Scholar
Morlighem, M., Seroussi, H., Larour, E., and Rignot, E. (2013b). Inversion of basal friction in Antarctica using exact and incomplete adjoints of a higher-order model. Journal of Geophysical Research: Earth Surface, 118(3), 1746–53. https://doi.org/10.1002/jgrf.20125.Google Scholar
Morlighem, M., Rignot, E., Mouginot, J. (2014). Deeply incised submarine glacial valleys beneath the Greenland Ice Sheet. Nature Geoscience, 7(6), 418–22. https://doi.org/10.1038/ngeo2167.Google Scholar
Morlighem, M., et al. (2017). BedMachine v3: Complete bed topography and ocean bathymetry mapping of Greenland from multi-beam echo sounding combined with mass conservation. Geophysical Research Letters, 44(21), 11,051–61. https://doi.org/10.1002/2017GL074954,2017GL074954.Google Scholar
Morlighem, M., et al. (2020). Deep glacial troughs and stabilizing ridges unveiled beneath the margins of the Antarctic ice sheet. Nature Geoscience, 13(2), 132–7. https://doi.org/10.1038/s41561-019-0510-8.Google Scholar
Nias, I. J., Cornford, S. L., and Payne, A. J. (2016). Contrasting the modelled sensitivity of the Amundsen Sea Embayment ice streams. Journal of Glaciology, 62(233), 552–62. https://doi.org/10.1017/jog.2016.40.Google Scholar
Nocedal, J., and Wright, S. (2006). Numerical Optimization. New York: Springer.Google Scholar
Pattyn, F. (2003). A new three-dimensional higher-order thermomechanical ice sheet model: Basic sensitivity, ice stream development, and ice flow across subglacial lakes. Journal of Geophysical Research: Solid Earth, 108(B8), 115. https://doi.org/10.1029/2002JB002329.Google Scholar
Payne, A. J., Vieli, A., Shepherd, A. P., Wingham, D. J., and Rignot, E. (2004). Recent dramatic thinning of largest West Antarctic ice stream triggered by oceans. Geophysical Research Letters, 31(23), 14. https://doi.org/10.1029/2004GL021284.Google Scholar
Perego, M., Price, S., and Stadler, G. (2014). Optimal initial conditions for coupling ice sheet models to Earth system models. Journal of Geophysical Research: Earth Surface, 119, 124. https://doi.org/10.1002/2014JF003181.Google Scholar
Petra, N., Zhu, H., Stadler, G., Hughes, T. J. R., and Ghattas, O. (2012). An inexact Gauss-Newton method for inversion of basal sliding and rheology parameters in a nonlinear Stokes ice sheet model. Journal of Glaciology, 58(211), 889903. https://doi.org/10.3189/2012JoG11J182.Google Scholar
Petra, N., Martin, J., Stadler, G., and Ghattas, O. (2014). A computational framework for infinitedimensional Bayesian inverse problems, Part II: Stochastic Newton MCMC with application to ice sheet flow inverse problems. SIAM Journal on Scientific Computing, 36(4), A1525A1555. https://doi.org/10.1137/130934805.Google Scholar
Pollard, D., and DeConto, R. M. (2012). A simple inverse method for the distribution of basal sliding coefficients under ice sheets, applied to Antarctica. The Cryosphere, 6(5), 953–71. https://doi.org/10.5194/tc-6–953–2012.Google Scholar
Pralong, M. R., and Gudmundsson, G. H. (2011). Bayesian estimation of basal conditions on Rutford Ice Stream, West Antarctica, from surface data. Journal of Glaciology, 57(202), 315–24.Google Scholar
Price, S. F., Payne, A. J., Howat, I. M., and Smith, B. E. (2011). Committed sea-level rise for the next century from Greenland ice sheet dynamics during the past decade. Proceedings of the National Academy of Sciences, USA, 108(22), 8978–83.Google Scholar
Quiquet, A., Dumas, C., Ritz, C., Peyaud, V., and Roche, D. M. (2018). The GRISLI ice sheet model (version 2.0): Calibration and validation for multi-millennial changes of the Antarctic ice sheet. Geoscientific Model Development, 11, 5003–25. https://doi.org/10.5194/gmd-11-5003-2018.Google Scholar
Ranganathan, M., Minchew, B. Meyer, C. R., and Gudmundsson, G. H. (2020). A new approach to inferring basal drag and ice rheology in ice streams, with applications to West Antarctic Ice Streams. Journal of Glaciology, 67(262), 229–42. https://doi.org/10.1017/jog.2020.95.Google Scholar
Rasmussen, L. A. (1988). Bed topography and mass-balance distribution of Columbia Glacier, Alaska, USA, determined from sequential aerial-photography. Journal of Glaciology, 34(117), 208–16.Google Scholar
Raymond, M. J., and Gudmundsson, G. H. (2009). Estimating basal properties of glaciers from surface measurements: a non-linear Bayesian inversion approach. The Cryosphere, 3(1), 181222.Google Scholar
Rignot, E., and Mouginot, J. (2012). Ice flow in Greenland for the International Polar Year 2008–2009. Geophysical Research Letters, 39(11). https://doi.org/10.1029/2012GL051634.Google Scholar
Rignot, E., Mouginot, J., and Scheuchl, B. (2011). Ice Flow of the Antarctic Ice Sheet. Science, 333(6048), 14271430. https://doi.org/10.1126/science.1208336.Google Scholar
Ritz, C., Edwards, T. L. Durand, G. et al. (2015). Potential sea-level rise from Antarctic ice-sheet instability constrained by observations. Nature, 528(7580), 115–18. https://doi.org/10.1038/nature16147.Google Scholar
Rommelaere, V., and MacAyeal, D. R. (1997). Large-scale rheology of the Ross Ice Shelf, Antarctica, computed by a control method. Annals of Glaciology, 24, 43–8.Google Scholar
Shapero, D., Badgeley, J., Hoffmann, A., and Joughin, I. (2021). icepack: A new glacier flow modeling package in Python, version 1.0. Geoscientific Model Development, 14, 4593–616. https://doi.org/10.5194/gmd-14-4593-2021.Google Scholar
Siegert, M., Ross, N., Corr, H., Kingslake, J., and Hindmarsh, R. (2013). Late Holocene ice-flow reconfiguration in the Weddell Sea sector of West Antarctica. Quaternary Science Reviews, 78, 98107.Google Scholar
Utke, J., Naumann, U., Fagan, M. (2008). OpenAD/F: A modular open-source tool for automatic differentiation of Fortran codes. ACM Transactions on Mathematical Software, 34(4) 136. https://doi.org/10.1145/1377596.1377598.Google Scholar
Waddington, E. D., Neumann, T. A., Koutnik, M. R., Marshall, H.-P., and Morse, D. L. (2007). Inference of accumulation-rate patterns from deep layers in glaciers and ice sheets. Journal of Glaciology, 53(183), 694712.Google Scholar
Warner, R., and Budd, W. (2000). Derivation of ice thickness and bedrock topography in data-gap regions over Antarctica, Annals of Glaciology, 31, 191–7.Google Scholar
Werder, M. A., Huss, M., Paul, F., Dehecq, A., and Farinotti, D. (2020). A Bayesian ice thickness estimation model for large-scale applications, Journal of Glaciology, 66(255), 137–52. https://doi.org/10.1017/jog. 2019.93.Google Scholar
Wernecke, A., Edwards, T. L., Nias, I. J., Holden, P. B. and Edwards, N. R.. (2020). Spatial probabilistic calibration of a high-resolution Amundsen Sea embayment ice sheet model with satellite altimeter data. The Cryosphere, 14(5), 1459–74.Google Scholar

References

Abdolghafoorian, A., and Farhadi, L. (2020). LIDA: A Land Integrated Data Assimilation framework for mapping land surface heat and evaporative fluxes by assimilating space-borne soil moisture and land surface temperature. Water Resources Research, 56(8), e2020WR027183.Google Scholar
Abrahart, R. J., Anctil, F., Coulibaly, P. et al. (2012). Two decades of anarchy? Emerging themes and outstanding challenges for neural network river forecasting. Progress in Physical Geography, 36(4), 480513.Google Scholar
Allen, G. H., Olden, J. D., Krabbenhoft, C. et al. (2020). Is our finger on the pulse? Assessing placement bias of the global river gauge network. AGU Fall Meeting Abstracts, H010-0016.Google Scholar
Altaf, M. U., El Gharamti, M., Heemink, A. W., and Hoteit, I. (2013). A reduced adjoint approach to variational data assimilation. Computer Methods in Applied Mechanics and Engineering, 254, 113.Google Scholar
Altenau, E. H., Pavelsky, T. M., Durand, M. T. et al. (2021). The Surface Water and Ocean Topography (SWOT) Mission River Database (SWORD): A global river network for satellite data products. Water Resources Research, 57(7), e2021WR030054.Google Scholar
Alvarez-Garreton, C., Ryu, D., Western, A. W. et al. (2015). Improving operational flood ensemble prediction by the assimilation of satellite soil moisture: Comparison between lumped and semi-distributed schemes. Hydrology and Earth System Sciences, 19(4), 1659–76.Google Scholar
Anderson, J. L. (2007). Exploring the need for localization in ensemble data assimilation using a hierarchical ensemble filter. Physica D: Nonlinear Phenomena, 230(1–2), 99111.Google Scholar
Andreadis, K. M., and Schumann, G. J. (2014). Estimating the impact of satellite observations on the predictability of large-scale hydraulic models. Advances in Water Resources, 73, 4454.Google Scholar
Annis, A., Nardi, F., and Castelli, F. (2022). Simultaneous assimilation of water levels from river gauges and satellite flood maps for near-real time flood mapping. Hydrology and Earth System Sciences, 26, 1019–41.Google Scholar
Annis, A., Nardi, F., Morrison, R. R., and Castelli, F. (2019). Investigating hydrogeomorphic floodplain mapping performance with varying DTM resolution and stream order. Hydrological Sciences Journal, 64(5), 525–38.Google Scholar
Arellano, L. N., Good, S. P., Sanchez-Murillo, R. et al. (2020). Bayesian estimates of the mean recharge elevations of water sources in the Central America region using stable water isotopes. Journal of Hydrology: Regional Studies, 32, 100739.Google Scholar
Avellaneda, P. M., Ficklin, D. L., Lowry, C. S., Knouft, J. H., and Hall, D. M. (2020). Improving hydrological models with the assimilation of crowdsourced data. Water Resources Research, 56, e2019WR026325.Google Scholar
Azimi, S., Dariane, A. B., Modanesi, S. et al. (2020). Assimilation of Sentinel 1 and SMAP – based satellite soil moisture retrievals into SWAT hydrological model: The impact of satellite revisit time and product spatial resolution on flood simulations in small basins. Journal of Hydrology, 581, 124367.Google Scholar
Babaeian, E., Sadeghi, M., Jones, S. B. et al. (2019). Ground, proximal, and satellite remote sensing of soil moisture. Review of Geophysics, 57(2), 530616.Google Scholar
Balsamo, G., Agusti-Panareda, A., Albergel, C. et al. (2018). Satellite and in situ observations for advancing global earth surface modelling: A review. Remote Sensing, 10, 38.Google Scholar
Bateni, S. M., Entekhabi, D., and Castelli, F. (2013). Mapping evaporation and estimation of surface control of evaporation using remotely sensed land surface temperature from a constellation of satellites. Water Resources Research, 49, 950–68.Google Scholar
Bateni, S., Entekhabi, D., Margulis, S., Castelli, F., and Kergoat, L. (2014). Coupled estimation of surface heat fluxes and vegetation dynamics from remotely sensed land surface temperature and fraction of photosynthetically active radiation. Water Resources Research, 50, 8420–40.Google Scholar
Bauser, H.H., Berg, D., and Roth, K. (2021). Technical Note: Sequential ensemble data assimilation in convergent and divergent systems. Hydrology and Earth System Sciences, 25, 3319–29.Google Scholar
Bauer-Marschallinger, B., Freeman, V., Cao, S. et al. (2019). Toward global soil moisture monitoring with Sentinel-1: Harnessing assets and overcoming obstacles. IEEE Transactions on Geoscience and Remote Sensing, 57(1), 520–39.Google Scholar
Baugh, C., de Rosnay, P., Lawrence, H. et al. (2020). The impact of SMOA soil moisture data assimilation within the operational global flood awareness system (GloFAS). Remote. Sensing, 12(9), 1490, https://doi.org/10.3390/rs12091490.Google Scholar
Berghuijs, W. R., Woods, R. A., Hutton, C. J., and Sivapalan, M. (2016). Dominant flood generating mechanisms across the United States. Geophysical Research Letters, 43(9), 4382–90.Google Scholar
Berthet, L., Andréassian, V., Perrin, C., and Javelle, P. (2009). How crucial is it to account for the antecedent moisture conditions in flood forecasting? Comparison of event-based and continuous approaches on 178 catchments. Hydrology and Earth System Sciences, 13(6), 819–31.Google Scholar
Beven, K., and Binley, A. (2014). GLUE: 20 years on. Hydrological Processes, 28(24), 5897–918.Google Scholar
Biancamaria, S., Lettenmaier, D. P., and Pavelsky, T. M. (2016). The SWOT mission and its capabilities for land hydrology. In Cazenave, A., Champollion, N., Benveniste, J., and Chen, J., eds., Remote Sensing and Water Resources. Cham: Springer, pp. 117–47.Google Scholar
Birkel, C., Soulsby, C., and Tetzlaff, D. (2014), Developing a consistent process-based conceptualization of catchment functioning using measurements of internal state variables. Water Resources Research, 50, 3481–501.Google Scholar
Bolten, J. D., Crow, W. T., Jackson, T. J., Zhan, X., and Reynolds, C. A. (2010). Evaluating the utility of remotely sensed soil moisture retrievals for operational agricultural drought monitoring. IEEE Journal of Selected Topics in Applied Earth Observations and Remote Sensing, 3(1), 5766.Google Scholar
Boucher, M.-A., Quilty, J., and Adamowski, J. (2020). Data assimilation for streamflow forecasting using extreme learning machines and multilayer perceptrons. Water Resources Research, 56, e2019WR026226. https://doi.org/10.1029/2019WR026226.Google Scholar
Boussetta, S., Balsamo, G., Beljaars, A. et al. (2013). Natural land carbon dioxide exchanges in the ECMWF Integrated Forecasting System: Implementation and offline validation, Journal of Geophysical Research: Atmospheres, 118, 5923–46.Google Scholar
Boussetta, S., Balsamo, G., Arduini, G. et al. (2021). ECLand: The ECMWF land surface modelling system. Atmosphere, 12, 723.Google Scholar
Burger, J., Le Brizaut, J. S., and Pogu, M. (1992). Comparison of two methods for the calculation of the gradient and of the Hessian of cost functions associated with differential systems. Mathematics and Computers in Simulation, 34, 551–62.Google Scholar
Caparrini, F., Castelli, F., and Entekhabi, D. (2003). Mapping of land-atmosphere heat fluxes and surface parameters with remote sensing data. Boundary-Layer Meteorology, 107, 605–33.Google Scholar
Caparrini, F., Castelli, F., and Entekhabi, D. (2004). Variational estimation of soil and vegetation turbulent transfer and heat flux parameters from sequences of multisensor imagery. Water Resources Research, 40, W12515.Google Scholar
Carrera, J., and Neuman, S.P. (1986). Estimation of aquifer parameters under transient and steady state conditions: 1. Maximum likelihood method incorporating prior information. Water Resources Research, 22(2), 199210.Google Scholar
Castillo, A., Castelli, F., and Entekhabi, D. (2015). Gravitational and capillary soil moisture dynamics for distributed hydrologic models. Hydrology and Earth System Sciences, 19(4), 1857–69.Google Scholar
Ceccatelli, M., Del Soldato, M., Solari, L. et al. (2021). Numerical modelling of land subsidence related to groundwater withdrawal in the Firenze-Prato-Pistoia basin (central Italy). Hydrogeology Journal, 29, 629–49.Google Scholar
Cenci, L., Pulvirenti, L., Boni, G. et al. (2017). An evaluation of the potential of Sentinel 1 for improving flash flood predictions via soil moisture-data assimilation. Advances in Geosciences, 44, 89100.Google Scholar
Chan, S. K., Bindlihs, R., O’Neill, P. et al. (2016). Assessment of the SMAP passive soil moisture product. IEEE Transactions on Geoscience and Remote Sensing, 54(8), 49945007.Google Scholar
Chan, S. K., Bindlihs, R., O’Neill, P. et al. (2017). Development and assessment of the SMAP enhanced passive soil moisture product. Remote Sensing of Environment, 204, 931–41.Google Scholar
Chen, J. M., and Liu, J. (2020). Evolution of evapotranspiration models using thermal and shortwave remote sensing data. Remote Sensing of Environment., 237, 111594.Google Scholar
Clark, M. P., Rupp, D. E., Woods, R. A. et al. (2008). Hydrological data assimilation with the ensemble Kalman filter: Use of streamflow observations to update states in a distributed hydrological model. Advances in Water Resources, 31(10), 1309–24.Google Scholar
Cooper, E. S., Dance, S. L., Garcia-Pintado, J., Nichols, N. K., and Smith, P. J. (2019). Observation operators for assimilation of satellite observations in fluvial inundation forecasting.Hydrology and Earth System Sciences, 23(6), 2541–59.Google Scholar
Das, N. N., Entekhabi, D., Dunbar, R. S. et al. (2018). The SMAP mission combined active‐passive soil moisture product at 9 km and 3 km spatial resolutions. Remote Sensing of Environment, 211, 204–17.Google Scholar
Dasgupta, A., Hostache, R., Ramsankaran, R. et al. (2021). On the impacts of observation location, timing and frequency on flood extent assimilation performance. Water Resources Research, 57, e2020WR028238Google Scholar
DeChant, C. M., and Moradkhani, H. (2012). Examining the effectiveness and robustness of sequential data assimilation methods for quantification of uncertainty in hydrologic forecasting. Water Resources Research, 48, W04518.Google Scholar
Domeneghetti, A., Castellarin, A., and Brath, A. (2012). Assessing rating-curve uncertainty and its effects on hydraulic model calibration. Hydrology and Earth System Sciences, 16(4), 1191–202.Google Scholar
Dong, J., Crow, W. T., Tobin, K. J. et al. (2020). Comparison of microwave remote sensing and land surface modeling for surface soil moisture climatology estimation. Remote Sensing of Environment., 242, 111756.Google Scholar
Emery, C. M., Biancamaria, S., Boone, A. et al. (2020). Assimilation of wide-swath altimetry water elevation anomalies to correct large-scale river routing model parameters.Hydrology and Earth System Sciences, 24, 2207–33.Google Scholar
Entekhabi, D., Nakamura, H., and Njoku, E.G. (1994). Solving the inverse problem for soil moisture and temperature profiles by sequential assimilation of multifrequency remotely sensed observations. IEEE Transactions on Geoscience and Remote Sensing, 32(2), 438–48.Google Scholar
Entekhabi, D., Rodriguez-Iturbe, I., and Castelli, F. (1996). Mutual interaction of soil moisture state and atmospheric processes. Journal of Hydrology, 184(1–2), 317.Google Scholar
Ercolani, G., and Castelli, F. (2017). Variational assimilation of streamflow data in distributed flood forecasting. Water Resources Research, 53(1), 158–83.Google Scholar
Fabryka-Martin, J., Merz, J., and Universities Council on Water Resources (1983). Hydrology: The Study of Water and Water Problems: A Challenge for Today and Tomorrow. Carbondale, IL: Universities Council on Water Resources.Google Scholar
Fan, Y. R., Huang, W. W., Li, Y. P., Huang, G. H., and Huang, K. (2015). A coupled ensemble filtering and probabilistic collocation approach for uncertainty quantification of hydrological models. Journal of Hydrology, 530, 255–72.Google Scholar
Frey, S. K., Miller, K., Khader, O. et al. (2021). Evaluating landscape influences on hydrologic behavior with a fully-integrated groundwater–surface water model. Journal of Hydrology, 602, 126758.Google Scholar
Garcia-Pintado, J., Mason, D. C., Dance, S. L. et al. (2015). Satellite-supported flood forecasting in river networks: A real case study. Journal of Hydrology, 523, 706–24.Google Scholar
Ghorbanidehno, H., Kokkinaki, A., Lee, J., and Darve, E. (2020). Recent developments in fast and scalable inverse modeling and data assimilation methods in hydrology. Journal of Hydrology, 591, 125266.Google Scholar
Gomez, B., Charlton-Perez, C. L., Lewis, H., and Candy, B. (2020). The Met Office operational soil moisture analysis system. Remote, 12, 3691.Google Scholar
Good, S. P., Noone, D., and Bowen, G. (2015), Hydrologic connectivity constrains partitioning of global terrestrial water fluxes. Science, 349(6244), 175–7.Google Scholar
Grimaldi, S., Li, Y., Pauwels, V. R. N., and Walker, J. P. (2016). Remote sensing-derived water extent and level to constrain hydraulic flood forecasting models: Opportunities and challenges. Surveys in Geophysics, 37(5), 9771034.Google Scholar
Hajj, M. E., Baghdadi, N., Zribi, M., and Bazzi, H. (2017). Synergic use of Sentinel-1 and Sentinel-2 images for operational soil moisture mapping at high spatial resolution over agricultural areas. Remote Sensing, 9(12), 1292. https://doi.org/10.3390/rs9121292.Google Scholar
Harrigan, S., Zsoter, E., Alfieri, L. et al. (2020). GloFAS-ERA5 operational global river discharge reanalysis 1979–present. Earth System Science Data, 12, 2043–60Google Scholar
Harvey, C. F., and Gorelick, S. M. (1995). Mapping hydraulic conductivity: Sequential conditioning with measurements of solute arrival time, hydraulic head, and local conductivity. Water Resources Research, 31(7), 1615–26.Google Scholar
Hochstetler, D. L., Barrash, W., Leven, C. et al. (2016). Hydraulic tomography: Continuity and discontinuity of high-K and low-K zones. Groundwater, 54(2), 171–85.Google Scholar
Hostache, R., Chini, M., Giustarini, L. et al. (2018). Near-real-time assimilation of SAR-derived flood maps for improving flood forecasts. Water Resources Research, 54(8), 5516–35.Google Scholar
Hou, Z.-Y., and Jin, Q.-N. (1997). Tikhonov regularization for nonlinear ill-posed problem. Nonlinear Analysis: Theory, Methods & Applications, 28(11), 1799–809.Google Scholar
Ishitsuka, Y., Gleason, C. J., Hagemann, M. W. et al. (2021). Combining optical remote sensing, McFLI discharge estimation, global hydrologic modeling, and data assimilation to improve daily discharge estimates across an entire large watershed. Water Resources Research, 56, e2020WR027794.Google Scholar
Kerr, Y. H., Waldteufel, P., Richaume, P. et al. (2012). The SMOS soil moisture retrieval algorithm. IEEE Transactions on Geoscience and Remote Sensing, 50(5), 1384–403.Google Scholar
King, F., Erler, A. R., Frey, S. K., and Flechter, C. G. (2020). Application of machine learning techniques for regional bias correction of snow water equivalent estimates in Ontario, Canada. Hydrology and Earth System Sciences, 24(10), 4887–902.Google Scholar
Kitanidis, P. K. (1996). On the geostatistical approach to the inverse problem. Advances in Water Resources. 19(6), 333–42.Google Scholar
Kolassa, J., Reichle, R. H., Liu, Q. et al. (2017). Data assimilation to extract soil moisture information from SMAP observations. Remote Sensing, 9, 1179.Google Scholar
Kustas, W. P., and Norman, J. M. (1999). Evaluation of soil and vegetation heat flux predictions using a simple two-source model with radiometric temperatures for partial canopy cover. Agriculture an Forest Meteorology, 94(1), 1329.Google Scholar
Laloy, E., Rogiers, B., Vrugt, J. A., Mallants, D., and Jacques, D. (2013). Efficient posterior exploration of a high-dimensional groundwater model from two-stage Markov chain Monte Carlo simulation and polynomial chaos expansion. Water Resources Research, 49(5), 2664–82.Google Scholar
Le Coz, J., Patalano, A., Collins, D. et al. (2016). Crowdsourced data for flood hydrology: Feedback from recent citizen science projects in Argentina, France and New Zealand. Journal of Hydrology, 541, 766–77.Google Scholar
Le Dimet, F. X., and Talagrand, O. (1986). Variational algorithms for analysis and assimilation of meteorological observations: theoretical aspects. Tellus A, 38:2, 97110.Google Scholar
Lee, J., Yoon, H., Kitanidis, P. K., Werth, C. J., and Valocchi, A. J. (2016). Scalable subsurface inverse modeling of huge data sets with an application to tracer concentration breakthrough data from magnetic resonance imaging, Water Resources Research, 52, 5213–31.Google Scholar
Lievens, H. et al. (2017). Joint Sentinel-1 and SMAP data assimilation to improve soil moisture estimates. Geophysical Research Letters, 44(12), 6145–53.Google Scholar
Liu, Y., Weerts, AH., Clark, M. et al. (2012). Advancing data assimilation in operational hydrologic forecasting: Progresses, challenges, and emerging opportunities. Hydrology and Earth System Sciences, 16(10), 3863–87.Google Scholar
Liu, K., Huang, G., Simunek, J. et al. (2021). Comparison of ensemble data assimilation methods for the estimation of time-varying soil hydraulic parameters. Journal of Hydrology, 594, 125729.Google Scholar
Maggioni, V., Vergara, H. J., Anagnostou, E. N. et al. (2013). Investigating the applicability of error correction ensembles of satellite rainfall products in river flow simulations. Journal of Hydrometeorology, 14(4), 1194–211.Google Scholar
Maliva, G. (2016). Aquifer Characterization Techniques. Cham Springer.Google Scholar
Man, J., Zheng, Q., Wu, L., and Zeng, L. (2020). Adaptive multi-fidelity probabilistic collocation-based Kalman filter for subsurface flow data assimilation: Numerical modeling and real-world experiment. Stochastic Environmental Research and Risk Assessment, 34, 1135–46. https://doi.org/10.1007/s00477-020-01815-yGoogle Scholar
Margulis, S. A., and Entekhabi, D. (2001). A coupled land surface-boundary layer model and its adjoint. Journal of Hydrometeorology, 2(3), 274–96.Google Scholar
Mason, D., Garcia-Pintado, J., Cloke, H. L., Dance, S. L., and Munoz-Sabatier, J. (2020) Assimilating high resolution remotely sensed data into a distributed hydrological model to improve run off prediction. ECMWF Tech Memo, 867, European Centre for Medium-Range Weather Forecasts.Google Scholar
Massari, C., Brocca, L., Tarpanelli, A., and Moramarco, T. (2015). Data assimilation of satellite soil moisture into rainfall-runoff modelling: A complex recipe? Remote Sensing, 7(9), 11403–33.Google Scholar
Massari, C., Camici, S., Ciabatta, L., and Brocca, L. (2018). Exploiting satellite-based surface soil moisture for flood forecasting in the Mediterranean area: State update versus rainfall correction. Remote Sensing, 10(2), 292. https://doi.org/10.3390/rs10020292.Google Scholar
Mazzoleni, M., Alfonso, L., and Solomatine, D.P. (2021). Exploring assimilation of crowdsourcing observations into flood models. In Scozzari, A., Mounce, S., Han, D., Soldovier, F., and Solomatine, D., eds., ICT for Smart Water Systems: Measurements and Data Science. Handbook of Environmental Chemistry, vol. 102. Cham: Springer, pp. 209–34.Google Scholar
Moges, E., Demissie, Y., Larsen, L., and Yassin, F. (2021). Review: Sources of hydrological model uncertainties and advances in their analysis. Water, 13(1), 28, https://doi.org/10.3390/w13010028.Google Scholar
Moradkhani, H., Hsu, K.-L., Gupta, H., and Sorooshian, S. (2005). Uncertainty assessment of hydrologic model states and parameters: Sequential data assimilation using the particle filter. Water Resources Research, 41, W05012.Google Scholar
Mudashiru, R. B., Sabtu, N., Abustan, I., and Balogun, W. (2021). Flood hazard mapping methods: A review. Journal of Hydrology, 603, 126846.Google Scholar
Nardi, F., Cudennec, C., Abrate, T. et al. (2021). Citizens AND HYdrology (CANDHY): conceptualizing a transdisciplinary framework for citizen science addressing hydrological challenges. Hydrological Sciences Journal. https://doi.org/10.1080/02626667.2020.1849707.Google Scholar
Neal, J. C., Atkinson, P. M., and Hutton, C. W. (2007). Flood inundation model updating using an ensemble Kalman filter and spatially distributed measurements. Journal of Hydrology, 336, 401–15.Google Scholar
Ng, L. W.-T., and Willcox, K. E. (2014). Multifidelity approaches for optimization under uncertainty. International Journal for Numerical Methods in Engineering, 100(10), 746–72.Google Scholar
Noilhan, J., and Planton, S. (1989). A simple parameterization of land-surface processes for meteorological models. Monthly Weather Review, 117, 536–50.Google Scholar
Ochoa-Rodriguez, S., Wang, L.-P., Gires, A. et al. (2015). Impact of spatial and temporal resolution of rainfall inputs on urban hydrodynamic modelling outputs: A multi-catchment investigation. Journal of Hydrology, 531(2), 389407.Google Scholar
Pierdicca, N., Chini, M., Pulvirenti, L. et al. (2009). Using COSMO-SkyMed data for flood mapping: Some case-studies. 2009 IEEE International Geoscience and Remote Sensing Symposium, 2, II-933–6.Google Scholar
Pinnington, E., Amezcua, J., Cooper, E. et al. (2021). Improving soil moisture prediction of a high-resolution land surface model by parameterising pedotransfer functions through assimilation of SMAP satellite data. Hydrology and Earth System Sciences, 25, 1617–41.Google Scholar
Rajabi, M. M., Ataie-Ashtiani, B., and Simmons, C. T. (2018). Model-data interaction in groundwater studies: Review of methods, applications and future directions. Journal of Hydrology, 567, 457–77.Google Scholar
Reichle, R. H., de Lannoy, G. J. M., Liu, Q. et al. (2017). Assessment of the SMAP Level-4 surface and root-zone soil moisture product using in situ measurements. Journal of Hydrometeorology, 18(10), 2621–45.Google Scholar
Reichle, R. H, Liu, Q., Koster, R. D. et al. (2019). Version 4 of the SMAP Level-4 soil moisture algorithm and data product. Journal of Advances in Modeling Earth Systems, 11(10), 3106–30.Google Scholar
Rodriguez-Fernandez, N., de Rosnay, P., Albergel, C. et al. (2019). SMOS neural network soil moisture data assimilation in a land surface model and atmospheric impact. Remote Sensing, 11, 1334.Google Scholar
Seneviratne, S. I., Corti, T., Davin, E. L. et al. (2010). Investigating soil moisture–climate interactions in a changing climate: A review. Earth-Science Reviews, 99(3–4), 125–61.Google Scholar
Sood, A., and Smakhtin, V. (2015). Global hydrological models: A review. Hydrological Sciences Journal, 60, 549–65.Google Scholar
Sprenger, M., Leistert, H., Gimbel, K., and Weiler, M. (2016a). Illuminating hydrological processes at the soil-vegetation-atmosphere interface with water stable isotopes. Reviews of Geophysics, 54(3), 674704.Google Scholar
Sprenger, M., Seeger, S., Blume, T., and Weiler, M. (2016b). Travel times in the vadose zone: Variability in space and time. Water Resources Research, 52(8), 5727–54.Google Scholar
Steklova, K., and Haber, E. (2017). Joint hydrogeophysical inversion: State estimation for seawater intrusion models in 3D. Computational Geosciences, 21, 7594.Google Scholar
Tada, M., Yoshimura, K., and Toride, K. (2021). Improving weather forecasting by assimilation of water vapor isotopes. Scientific Reports, 11(1), 18067.Google Scholar
Talagrand, O., and Courtier, P. (1987). Variational assimilation of meteorological observations with the adjoint vorticity equation. I: Theory. Quarterly Journal of the Royal Meteorological Society, 113(478), 1311–28.Google Scholar
Thacker, W. C. (1989). The role of the Hessian matrix in fitting models to measurements. Journal of Geophysical Research: Oceans, 94(C5), 6177–96.Google Scholar
Verstraeten, W. W., Veroustraete, F., van der Sande, C. J., Grootaers, I., and Feyen, J. (2006). Soil moisture retrieval using thermal inertia, determined with visible and thermal spaceborne data, validated for European forests. Remote Sensing of Environment, 101(3), 299314.Google Scholar
de Vos, L. W., Droste, A .M., Zander, M. J. et al. (2020). Hydrometeorological monitoring using opportunistic sensing networks in the Amsterdam metropolitan area. Bulletin of the American Meteorological Society, 101(2), E167E185.Google Scholar
Wagner, W., Hahn, S., Kidd, R. et al. (2013). The ASCAT soil moisture product: A review of its specifications, validation results, and emerging applications. Meteorologische Zeitschrift, 22(1), 533.Google Scholar
Waller, J. A., Garcia-Pintado, J., Mason, D. C., Dance, S. L., and Nichols, N. K. (2018) Technical note: Assessment of observation quality for data assimilation in flood models. Hydrology and Earth System Sciences, 22(7), 3983–92.Google Scholar
Wang, K., and Dickinson, R. E. (2012). A review of global terrestrial evapotranspiration: Observation, modeling, climatology, and climatic variability. Reviews of Geophysics, 50(2), RG2005.Google Scholar
Wang, L. L., Chen, D. H., and Bao, H. J. (2016). The improved Noah land surface model based on storage capacity curve and Muskingum method and application in GRAPES model. Atmospheric Science Letters, 17, 190–8.Google Scholar
Weeser, B., Stenfert Kroese, J., Jacobs, S. R. et al. (2018). Citizen science pioneers in Kenya: A crowdsourced approach for hydrological monitoring. Science of the Total Environment, 631–632, 1590–9.Google Scholar
Xie, X., Meng, S., Liang, S., and Yao, Y. (2014). Improving streamflow predictions at ungauged locations with real-time updating: Application of an EnKF-based state-parameter estimation strategy. Hydrology and Earth System Sciences, 18(10), 3923–36.Google Scholar
Xu, T., Bateni, S. M., Liang, S., Entekhabi, D., and Mao, K. (2014). Estimation of surface turbulent heat fluxes via variational assimilation of sequences of land surface temperatures from Geostationary Operational Environmental Satellites. Journal of Geophysical Research: Atmospheres, 119(18), 10780–98.Google Scholar
Xu, T., He, X., Bateni, S. M. et al. (2019). Mapping regional turbulent heat fluxes via variational assimilation of land surface temperature data from polar orbiting satellites. Remote Sensing of Environment, 221, 444–61.Google Scholar
Yang, K., Zhu, L., Chen, Y. et al. (2016). Land surface model calibration through microwave data assimilation for improving soil moisture simulations. Journal of Hydrology, 533, 266–76.Google Scholar
Zeng, X., and Decker, M. (2009). Improving the numerical solution of soil moisture–based Richards equation for land models with a deep or shallow water table. Journal of Hydrometerology., 1081, 308–19.Google Scholar
Zhang, H., Franssen, H.-J. H., Han, X., Vrugt, J. A., and Vereecken, H. (2017). State and parameter estimation of two land surface models using the ensemble Kalman filter and the particle filter. Hydrology and Earth System Sciences, 21(9), 4927–58.Google Scholar
Zhang, F., Zhang, L. W., Shi, J. J., and Huang, J. F. (2014). Soil moisture monitoring based on land surface temperature‐vegetation index space derived from MODIS data. Pedosphere, 24(4), 450–60.Google Scholar
Zheng, W., Zhan, X., Liu, J. and Ek, M. (2018). A preliminary assessment of the impact of assimilating satellite soil moisture data products on NCEP Global Forecast System. Advances in Meteorology, 1–12, 7363194.Google Scholar
Zsoter, E., Cloke, H., Stephens, E. et al. (2019). How well do operational numerical weather prediction configurations represent hydrology? Journal of Hydrometeorology, 20(8), 1533–52.Google Scholar

References

Anderson, J. L. (2001). An ensemble adjustment Kalman filter for data assimilation. Monthly Weather Review, 129, 2884–903.Google Scholar
Arellano, A. F., Jr., Kasibhatla, P. S., Giglio, L. et al. (2006). Time dependent inversion estimates of global biomass-burning CO emissions using measurement of pollution in the troposphere (MOPITT) measurements. Journal of Geophysical Research, 111. https://doi.org/10.1029/2005JD006613.Google Scholar
Arellano, A. F., Jr., Raeder, K., Anderson, J. L. et al. (2007). Evaluating model performance of an ensemble-based chemical data assimilation system during INTEX-B field mission. Atmospheric Chemistry and Physics, 7, 5695–710.Google Scholar
Barkley, M. P., Smedt, I. D., Roozendael, M. V. et al. (2013). Top-down isoprene emissions over tropical South America inferred from SCIAMACHY and OMI formaldehyde columns. Journal of Geophysical Research: Atmospheres, 118(12), 6849–68. https://doi.org/10.1002/jgrd.50552.Google Scholar
Barré, J., Gaubert, B., Arellano, A. F. J. et al. (2015). Assessing the impacts of assimilating IASI and MOPITT CO retrievals using CESMCAM-chem and DART. Journal of Geophysical Research, 120, 10501–29. https://doi.org/10.1002/2015JD023467.Google Scholar
Basu, S., Guerlet, S., Butz, A. et al. (2013). Global CO2 fluxes estimated from GOSAT retrievals of total column CO2. Atmospheric Chemistry and Physics, 13, 8695–717. https://doi.org/10.5194/acp-13-8695-2013.Google Scholar
Bauwens, M., Stavrakou, T., Muller, J.-F. et al. (2016). Nine years of global hydrocarbon emissions based on source inversion of OMI formaldehyde observations. Atmospheric Chemistry and Physics, 16, 10133–58. https://doi.org/10.5194/acp-16-10133-2016.Google Scholar
Byrne, B., Jones, D. B. A., Strong, K. et al. (2017). Sensitivity of CO2 surface flux constraints to observational coverage. Journal of Geophysical Research, 122, 6672–94. https://doi.org/10.1002/2016JD026164.Google Scholar
Crawford, J. H., Heald, C. L., Fuelberg, H. E. et al. (2004). Relationship between Measurements of Pollution in the Troposphere (MOPITT) and in situ observations of CO based on a large-scale feature sampled during TRACE-P. Journal of Geophysical Research, 109, D15S04. https://doi.org/10.1029/2003JD004308.Google Scholar
Crowell, S., Baker, D., Schuh, A. et al. (2019). The 2015–2016 carbon cycle as seen from OCO-2 and the global in situ network. Atmospheric Chemistry and Physics, 19, 9797–831. https://doi.org/10.5194/acp-19-9797-2019.Google Scholar
Deeter, M. N., Worden, H. M., Edwards, D. P., Gille, J. C., and Andrews, A. E. (2012). Evaluation of MOPITT retrievals of lower-tropospheric carbon monoxide over the United States. Journal of Geophysical Research, 117 (D13306). https://doi.org/10.1029/2012JD017553.Google Scholar
Deng, F., Jones, D., O’Dell, C. W., Nassar, R., and Parazoo, N. C. (2016). Combining GOSAT XCO2 observations over land and ocean to improve regional CO2 flux estimates. Journal of Geophysical Research, 121, 1896–913. https://doi.org/10.1002/2015JD024157.Google Scholar
Derber, J. C. (1989). A variational continuous assimilation technique. Monthly Weather Review, 117, 2437–46. https://doi.org/10.1175/1520-0493(1989)117<2437:AVCAT>2.0.CO;2.Google Scholar
Elbern, H., and Schmidt, H. (2001). Ozone episode analysis by four-dimensional variational chemistry data assimilation. Journal of Geophysical Research, 106 (D4), 3569–90.Google Scholar
Elguindi, N., Granier, C., Stavrakou, T. et al. (2020). Intercomparison of magnitudes and trends in anthropogenic surface emissions from bottom-up inventories, top-down estimates, and emission scenarios. Earth’s Future, 8 (e2020EF001520). https://doi.org/10.1029/2020EF001520.Google Scholar
Enting, I. G., Trudinger, C. M., and Francey, R. J. (1995). A synthesis inversion of the concentration and δ13C of atmospheric CO2. Tellus, 47B, 3552.Google Scholar
Feng, L., Palmer, P. I., Parker, R. J. et al. (2016). Estimates of European uptake of CO2 inferred from GOSAT XCO2 retrievals: Sensitivity to measurement bias inside and outside Europe. Atmospheric Chemistry and Physics, 16, 1289–302. https://doi.org/10.5194/acp-16-1289-2016.Google Scholar
Feng, L., Palmer, P. I., Yang, Y. et al. (2011). Evaluating a 3-D transport model of atmospheric CO2 using ground-based, aircraft, and space-borne data. Atmospheric Chemistry and Physics, 11, 2789–803. https://doi.org/10.5194/acp-11-2789-.Google Scholar
Gaubert, B., Emmons, L. K., Raeder, K. et al. (2020). Correcting model biases of CO in East Asia: Impact on oxidant distributions during KORUS-AQ. Atmospheric Chemistry and Physics, 20, 14617–47. https://doi.org/10.5194/acp20-.Google Scholar
GBD 2019 Risk Factors Collaborators. (2020). Global burden of 87 risk factors in 204 countries and territories, 1990–2019: A systematic analysis for the global burden of disease study 2019. Lancet, 396, 1223–49. https://doi.org/10.1016/S0140-6736(20)30752-2.Google Scholar
Hamill, T. M., Whitaker, J. S., and Snyder, C. (2001). Distance-dependent filtering of background error covariance estimates in an ensemble Kalman filter. Monthly Weather Review, 129, 2776–90.Google Scholar
Heald, C. L., Jacob, D. J., Jones, D. B. A. et al. (2004). Comparative inverse analysis of satellite (MOPITT) and aircraft (TRACE-P) observations to estimate Asian sources of carbon monoxide. Journal of Geophysical Research, 109 (D23306). https://doi.org/10.1029/2004JD005185.Google Scholar
Hooghiemstra, P. B., Krol, M. C., Bergamaschi, P. et al. (2012). Comparing optimized CO emission estimates using MOPITT or NOAA surface network observations. Journal of Geophysical Research, 117 (D06309). https://doi.org/https://doi.org/10.1029/2011JD017043.Google Scholar
Hunt, B. R., Kostelich, E. J., and Szunyogh, I. (2007). Efficient data assimilation for spatiotemporal chaos: A local ensemble transform Kalman filter. Physica D, 230, 112–26.Google Scholar
Jacobson, A. R., Schuldt, K. N., Miller, J. B. et al. (2020). CarbonTracker CT2019B. https://doi.org/10.25925/20201008.Google Scholar
Jiang, Z., Jones, D. B. A., Worden, H. M., and Henze, D. K. (2015). Sensitivity of top-down CO source estimates to the modeled vertical structure in atmospheric CO. Atmospheric Chemistry and Physics, 15, 1521–37. https://doi.org/10.5194/acp-15-1521-2015.Google Scholar
Jiang, Z., Worden, J. R., Worden, H. et al. (2017). A 15-year record of CO emissions constrained by MOPITT CO observations. Atmospheric Chemistry and Physics, 17, 4565–83. https://doi.org/10.5194/acp-17-4565-2017.Google Scholar
Jiang, Z., Zhu, R., Miyazaki, K. et al. (2022). Decadal variabilities in tropospheric nitrogen oxides over United States, Europe, and China. Journal of Geophysical Research: Atmospheres, 127. https://doi.org/10.1029/2021JD035872.Google Scholar
Jones, D. B. A., Bowman, K. W., Logan, J. A. et al. (2009). The zonal structure of tropical O3 and CO as observed by the Tropospheric Emission Spectrometer in 2004. Part 1: Inverse modeling of CO emissions. Atmospheric Chemistry and Physics, 9, 3547–62.Google Scholar
Khattatov, B. V., Gille, J. C., Lyjak, L. V. et al. (1999). Assimilation of photochemically active species and a case analysis of UARS data. Journal of Geophysical Research (D15), 18715–37.Google Scholar
Kopacz, M., Jacob, D. J., Fisher, J. A. et al. (2010). Global estimates of CO sources with high resolution by adjoint inversion of multiple satellite datasets (MOPITT, AIRS, SCIAMACHY, TES). Atmospheric Chemistry and Physics, 10, 855–76.Google Scholar
Le Quéré, C., Andrew, R. M., Friedlingstein, P. et al. (2018). Global Carbon Budget 2017. Earth System Science Data, 405–48. https://doi.org/10.5194/essd10–405–2018.Google Scholar
Levelt, P. F., Khattatov, B. V., Gille, J. C. et al. (1998). Assimilation of MLS ozone measurements in the global three-dimensional chemistry transport model ROSE. Geophysical Research Letters, 25(24), 4493–96.Google Scholar
Liu, J., Bowman, K., Lee, M. et al. (2014). Carbon monitoring system flux estimation and attribution: Impact of ACOS-GOSAT XCO2 sampling on the inference of terrestrial biospheric sources and sinks. Tellus B, 66, 22486. https://doi.org/10.3402/tellusb.v66.22486.Google Scholar
Liu, J., Bowman, K. W., and Henze, D. K. (2015). Source-receptor relationships of column average CO2 and implications for the impact of observations on flux inversions. Journal of Geophysical Research, 120, 5214–36. https://doi.org/10.1002/2014JD022914.Google Scholar
Maasakkers, J. D., Jacob, D. J., Sulprizio, M. P. et al. (2019). Global distribution of methane emissions, emission trends, and OH concentrations and trends inferred from an inversion of GOSAT satellite data for 2010–2015. Atmospheric Chemistry and Physics, 19, 7859–81. https://doi.org/10.5194/acp-19-7859-2019.Google Scholar
Marais, E. A., Jacob, D. J., Kurosu, T. P. et al. (2012). Isoprene emissions in Africa inferred from OMI observations of formaldehyde columns. Atmospheric Chemistry and Physics, 12(14), 6219–35. https://doi.org/10.5194/acp-12–6219–2012Google Scholar
Millet, D. B., Jacob, D. J., Turquety, S. et al. (2006). Formaldehyde distribution over North America: Implications for satellite retrievals of formaldehyde columns and isoprene emission. Journal of Geophysical Research: Atmospheres, 111(D24). https://doi.org/10.1029/2005JD006853.Google Scholar
Miyazaki, K., Bowman, K., Sekiya, T. et al. (2020). Updated tropospheric chemistry reanalysis and emission estimates, TCR-2, for 2005–2018. Earth System Science Data, 12, 2223–59. https://doi.org/10.5194/essd-12-2223-2020.Google Scholar
Miyazaki, K., Eskes, H., Sudo, K. et al. (2017). Decadal changes in global surface NOx emissions from multi-constituent satellite data assimilation. Atmospheric Chemistry and Physics, 17, 807–37. https://doi.org/10.5194/acp-17-807-2017.Google Scholar
Miyazaki, K., Eskes, H. J., Sudo, K. (2012). Simultaneous assimilation of satellite NO2, O3, CO, and HNO3 data for the analysis of tropospheric chemical composition and emissions. Atmospheric Chemistry and Physics, 12(20), 9545–79. https://doi.org/10.5194/acp-12-9545-2012.Google Scholar
Müller, J.-F., and Stavrakou, T. (2005). Inversion of co and NOX emissions using the adjoint of the images model. Atmospheric Chemistry and Physics, 5(5), 1157–86. https://doi.org/10.5194/acp-5-1157-2005.Google Scholar
Palmer, P. I., Feng, L., Baker, D. et al. (2019). Net carbon emissions from African biosphere dominate pan-tropical atmospheric CO2 signal. Nature Communications, 10, 3344. https://doi.org/10.1038/s41467-019-11097-w.Google Scholar
Palmer, P. I., Jacob, D. J., Fiore, A. M. eet al. (2003). Mapping isoprene emissions over North America using formaldehyde column observations from space. Journal of Geophysical Research: Atmospheres, 108 (D6). https://doi.org/10.1029/2002JD002153.Google Scholar
Parrington, M., Jones, D. B. A. Bowman, K. W. et al. (2008). Estimating the summertime tropospheric ozone distribution over North America through assimilation of observations from the tropospheric emission spectrometer. Journal of Geophysical Research, 113 (D18307). https://doi.org/10.1029/2007JD009341.Google Scholar
Pétron, G., Granier, C., Khattotov, B. et al. (2004). Monthly CO surface sources inventory based on the 2000–2001 MOPITT satellite data. Geophysical Research Letters, 31 (L21107). https://doi.org/10.1029/2004GL020560.Google Scholar
Qu, Z., Henze, D. K., Worden, H. M. et al. (2022). Sector-based top-down estimates of NOx, SO2, and CO emissions in East Asia. Geophysical Research Letters, 49 (e2021GL096009). https://doi.org/10.1029/2021GL096009.Google Scholar
Rabier, F., J¨arvinen, H., Klinker, E., Mahfouf, J.-F., and Simmons, A. (2000). The ECMWF operational implementation of four-dimensional variational assimilation. I: Experimental results with simplified physics. Quarterly Journal of the Royal Meteorological Society, 126, 1143–70.Google Scholar
Rigby, M., Montzka, S. A., Prine, R. G. et al. (2017). Role of atmospheric oxidation in recent methane growth. Proceedings of the National Academy of Sciences, 114(21), 5373–7. https://doi.org/10.1073/pnas.1616426114.Google Scholar
Riishøjgaard, L. P., Štajner, I., and Lou, G.-P. (2000). The GEOS ozone data assimilation system. Advances in Space Research, 25, 1063–72. https://doi.org/10.1016/S0273-1177(99)00443-.Google Scholar
Rodgers, C. D. (2000). Inverse Methods for Atmospheric Sounding: Theory and Practice. Singapore: World Scientific Publishing.Google Scholar
Rodgers, C. D., and Connor, B. J. (2003). Intercomparison of remote sounding instruments. Journal of Geophysical Research, 108(D3), 4116. https://doi.org/10.1029/2002jd002299.Google Scholar
Sitch, S., Friedlingstein, P., Gruber, N. et al. (2015). Recent trends and drivers of regional sources and sinks of carbon dioxide. Biogeosciences, 12, 653–79. https://doi.org/10.5194/bg-12-653-2015.Google Scholar
Stanevich, I., Jones, D. B. A., Strong, K. et al. (2021). Characterizing model errors in chemical transport modeling of methane: Using GOSAT XCH4 data with weak-constraint four-dimensional variational data assimilation. Atmospheric Chemistry and Physics, 21, 9545–72. https://doi.org/10.5194/acp-21-9545-2021.Google Scholar
Stavrakou, T., and Müller, J.-F. (2006). Grid-based versus big region approach for inverting CO emissions using measurement of pollution in the troposphere (MOPITT) data. Journal of Geophysical Research, 111(D15), 304. https://doi.org/10.1029/2005JD006896.Google Scholar
Trémolet, Y. (2006). Accounting for an imperfect model in 4d-var. Quarterly Journal of the Royal Meteorological Society, 132, 2483–504. https://doi.org/10.1256/qj.05.224.Google Scholar
Trémolet, Y. (2007). Model-error estimation in 4d-var. Quarterly Journal of the Royal Meteorological Society, 133, 1267–80. https://doi.org/10.1002/qj.94.Google Scholar
Turner, A. J., Frankenberg, C., and Kort, E. A. (2019). Interpreting contemporary trends in atmospheric methane. Proceedings of the National Academy of Sciences, 116(8), 2805–13. https://doi.org/10.1073/pnas.1814297116.Google Scholar
Worden, H. M., Deeter, M. N., Edwards, D. P. et al. (2010). Observations of near-surface carbon monoxide from space using MOPITT multispectral retrievals. Journal of Geophysical Research, 115,(D18), 314. https://doi.org/10.1029/2010JD014242.Google Scholar
Yoshida, Y., Ota, Y., Eguchi, N. et al. (2011). Retrieval algorithm for CO2 and CH4 column abundances from short-wavelength infrared spectral observations by the greenhouse gases observing satellite. Atmospheric Measurement Techniques, 4, 717–34. https://doi.org/10.5194/amt-4-717-2011.Google Scholar
Zhang, X., Jones, D. B. A., Keller, M. et al. (2019). Quantifying emissions of co and nox using observations from MOPITT, OMI, TES, and OSIRIS. Journal of Geophysical Research, 124 (1029). https://doi.org/11701193/2018JD028670.Google Scholar
Zheng, B., Chevallier, F., Yin, Y. et al. (2019). Global atmospheric carbon monoxide budget 2000–2017 inferred from multi-species atmospheric inversions. Earth System Science Data, 11, 1411–36. https://doi.org/10.5194/essd-11-1411-2019.Google Scholar

References

Amezcua, J., and Van Leeuwen, P. J. (2014). Gaussian anamorphosis in the analysis step of the EnKF: A joint state-variable/observation approach. Tellus A: Dynamic Meteorology and Oceanography, 6(1), 23493. http://doi.org/10.3402/tellusa.v66.23493.Google Scholar
Anderson, J. L., and Anderson, S. L. (1999). A Monte Carlo implementation of the nonlinear filtering problem to produce ensemble assimilations and forecasts. Monthly Weather Review, 127(12), 2741–58.Google Scholar
Balis, D., Koukouli, M. E., Siomos, N. et al. (2016). Validation of ash optical depth and layer height retrieved from passive satellite sensors using EARLINET and airborne lidar data: The case of the Eyjafjallajokull eruption. Atmospheric Chemistry and Physics, 16, 5705–20.Google Scholar
Bishop, C. H., Brian, J. E., and Sharanya, J. M. (2001). Adaptive sampling with the Ensemble Transform Kalman Filter. Part 1: Theoretical aspects. Monthly Weather Review, 129(3), 420–36.Google Scholar
Bishop, C. H. (2016). The GIGG-EnKF: Ensemble Kalman filtering for highly skewed non-negative uncertainty distributions. Quarterly Journal of the Royal Meteorological Society, 142(696), 1395–412.Google Scholar
Boichu, M., Clarisse, L., Khvorostyanov, D., and Clerbaux, C. (2014). Improving volcanic sulfur dioxide cloud dispersal forecasts by progressive assimilation of satellite observations, Geophysical Research Letters. American Geophysical Union, 41, 2637–43.Google Scholar
Bonadonna, C., Folch, A., Loughlin, S., and Puempel, H. (2012). Future developments in modelling and monitoring of volcanic ash clouds: Outcomes from the first IAVCEI-WMO workshop on ash dispersal forecast and civil aviation. Bulletin of Volcanology, 74, 110.Google Scholar
Burgers, G., Leeuwen, P. J. van, and Evensen, G. (1998). Analysis scheme in the ensemble Kalman filter. Monthly Weather Review, 126(6), 1719–24.Google Scholar
Carrassi, A., Bocquet, M., Bertino, L., and Evensen, G. (2018). Data assimilation in the geosciences: An overview of methods, issues, and perspectives. WIREs Climate Change, 9(e535), 150.Google Scholar
Carn, S. A., Krueger, A. J., Krotkov, N. A., Yang, K., and Evans, K. (2009). Tracking volcanic sulfur dioxide clouds for aviation hazard mitigation. Natural Hazards, 51, 325–43.Google Scholar
Chai, T., Crawford, A., Stunder, B. et al. (2017). Improving volcanic ash predictions with the HYSPLIT dispersion model by assimilating MODIS satellite retrievals. Atmospheric Chemistry and Physics, 17, 2865–79.Google Scholar
Clarisse, L., Hurtmans, D., Clerbaux, C. et al. (2012). Retrieval of sulphur dioxide from the infrared atmospheric sounding interferometer (IASI). Atmospheric Measurement Techniques, 5, 581–94.Google Scholar
Clarkson, R., and Simpson, H. (2017). Maximising airspace use during volcanic eruptions: Matching engine durability against ash cloud occurrence. In Proceedings of the NATO STO AVT-272 Specialists Meeting on Impact of Volcanic Ash Clouds on Military Operations, vol. 1, pp. 17-1–17-19. www.sto.nato.int/publications/STO%20Meeting%20Proceedings/STO-MP-AVT-272/MP-AVT-272-17.pdf.Google Scholar
Costa, A., Suzuki, Y., Cerminara, M. et al. (2016). Results of the eruptive column model inter-comparison study, Journal of Volcanology and Geothermal Research, 326, 225.Google Scholar
Crawford, A., Stunder, B., Ngan, F., and Pavalonis, M. (2016). Initializing HYSPLIT with satellite observations of volcanic ash: A case study of the 2008 Kasatochi eruption. Journal of Geophysical Research: Atmospheres, 121(10), 786–10.Google Scholar
Dunn, M. G., and Wade, D. P. (1994). Influence of volcanic ash clouds on gas turbine engines. In Casadevall, T. J., ed., Volcanic ash and aviation safety; Proceedings of the First International Symposium on Volcanic Ash and Aviation Safety held in Seattle, Washington, in July 1991. Reston, VA: US Geological Survey Bulletin 2047, pp. 107–17. https://doi.org/10.3133/b2047.Google Scholar
Eckhardt, S., Prata, A. J., Seibert, P., Stebel, K., and Stohl, A. (2008). Estimation of the vertical profile of sulfur dioxide injection into the atmosphere by a volcanic eruption using satellite column measurements and inverse transport modeling. Atmospheric Chemistry and Physics, 8, 3881–97.Google Scholar
Eliasson, J., and Yoshitani, J. (2015). Airborne measurements of volcanic ash and current state of ash cloud prediction, Disaster Prevention Research Institute Annuals, 58B, 3541. www.dpri.kyoto-u.ac.jp/nenpo/no58/ronbunB/a58b0p03.pdf.Google Scholar
Evensen, G. (1994). Sequential data assimilation with a nonlinear quasi-geostrophic model using Monte Carlo methods to forecast error statistics. Journal of Geophysical Research: Oceans, 99(C5), 10143–62.Google Scholar
Evensen, G. (2003). The ensemble Kalman filter: Theoretical formulation and practical implementation. Ocean Dynamics, 53(4), 343–67.Google Scholar
Flemming, J., and Inness, A. (2013). Volcanic sulfur dioxide plume forecasts based on UV satellite retrievals for the 2011 Grímsvötn and the 2010 Eyjafjallajökull eruption. Journal of Geophysical Research: Atmospheres, 118(17), 10172–89.Google Scholar
Folch, A. (2012). A review of tephra transport and dispersal models: Evolution, current status, and future perspectives. Journal of Volcanology and Geothermal Research, 235–236, 96115.Google Scholar
Folch, A., Mingari, L., Gutierrez, N. et al. (2020). FALL3D-8.0: a computational model for atmospheric transport and deposition of particles, aerosols and radionuclides. Part 1: Model physics and numerics. Geoscientific Model Development, 13, 1431–58.Google Scholar
Francis, P. N., Cooke, M. C., and Saunders, R.W. (2012). Retrieval of physical properties of volcanic ash using Meteosat: A case study from the 2010 Eyjafjallajökull eruption. Journal of Geophysical Research: Atmospheres, 117, D00U09. https://doi.org/10.1029/2011JD016788Google Scholar
Fu, G., Lin, H.X., Heemink, A.W. et al. (2015). Assimilating aircraft-based measurements to improve forecast accuracy of volcanic ash transport. Atmospheric Environment, 115, 170–84.Google Scholar
Fu, G., Heemink, A., Lu, S. et al. (2016). Model-based aviation advice on distal volcanic ash clouds by assimilating aircraft in situ measurements. Atmospheric Chemistry and Physics, 16(14), 9189–200.Google Scholar
Fu, G., Prata, F., Lin, H. X. et al. (2017). Data assimilation for volcanic ash plumes using a satellite observational operator: A case study on the 2010 Eyjafjallajökull volcanic eruption. Atmospheric Chemistry and Physics, 17(2), 1187–205.Google Scholar
Houtekamer, P. L., and Zhang, F. (2016). Review of the ensemble Kalman filter for atmospheric data assimilation. Monthly Weather Review, 144(12), 4489–532.Google Scholar
Hunt, B. R., Kostelich, E. J., and Szunyogh, I. (2007). Efficient data assimilation for spatiotemporal chaos: A local ensemble transform Kalman filter. Physica D, 230(1), 112–26.Google Scholar
Inness, A., Ades, M., Balis, D. et al. (2022). The CAMS volcanic forecasting system utilizing near-real time data assimilation of S5P/TROPOMI SO2 retrievals. Geoscientific Model Development, 15, 971–94.Google Scholar
Kalman, R. E. (1960). A new approach to linear filtering and prediction problems. Journal of Basic Engineering, 82, 3545.Google Scholar
Kristiansen, N. I., Stohl, A., Prata, A. J. et al. (2010). Remote sensing and inverse transport modeling of the Kasatochi eruption sulfur dioxide cloud. Journal of Geophysical Research: Atmospheres, 115, 118.Google Scholar
Kristiansen, N., Stohl, A., Prata, A. et al. (2012). Performance assessment of a volcanic ash transport model mini-ensemble used for inverse modeling of the 2010 Eyjafjallajökull eruption. Journal of Geophysical Research: Atmospheres, 117, 125.Google Scholar
Lu, S., Lin, H. X., Heemink, A. W., Fu, G., and Segers, A. J. (2016a). Estimation of volcanic ash emissions using trajectory-based 4D-Var data assimilation. Monthly Weather Review, 144(2), 575–89.Google Scholar
Lu, S., Lin, H. X., Heemink, A., Segers, A., and Fu, G. (2016b). Estimation of volcanic ash emissions through assimilating satellite data and ground-based observations. Journal of Geophysical Research: Atmospheres, 121(18), 10971–94.Google Scholar
Mastin, L. G., Guffanti, M. C., and Servranckx, R. et al. (2009). A multidisciplinary effort to assign realistic source parameters to models of volcanic ash-cloud transport and dispersion during eruptions. Journal of Volcanology and Geothermal Research, 186(1–2), 1021.Google Scholar
Miller, T. P., and Casadevall, T. J. (2000). Volcanic ash hazards to aviation. In Sigurdsson, H., ed., Encyclopedia of Volcanoes.San Diego, CA: Academic Press, pp. 915–30.Google Scholar
Mingari, L., Folch, A., Prata, A. T. et al. (2022). Data assimilation of volcanic aerosols using FALL3D+PDAF. Atmospheric Chemistry and Physics, 22, 1773–92.Google Scholar
Moxnes, E. D., Kristiansen, N. I., Stohl, A. et al. (2014). Separation of ash and sulfur dioxide during the 2011 Grímsvötn eruption. Journal of Geophysical Research: Atmospheres, 119, 7477–01.Google Scholar
Muser, L. O., Hoshyaripour, G. A., Bruckert, J. et al. (2020). Particle aging and aerosol–radiation interaction affect volcanic plume dispersion: Evidence from the Raikoke 2019 eruption. Atmospheric Chemistry and Physics, 20, 15015–36.Google Scholar
Nerger, L., Janjić, T., Schröter, J., and Hiller, W. (2012). A unification of ensemble square root Kalman filters. Monthly Weather Review, 140(7), 2335–45.Google Scholar
Osores, S., Ruiz, J., Folch, A., and Collini, E. (2020). Volcanic ash forecast using ensemble-based data assimilation: An ensemble transform Kalman filter coupled with the Fall3d-7.2 Model (ETKF–Fall3d Version 1.0). Geoscientific Model Development, 13(1), 122.Google Scholar
Pardini, F., Corradini, S., Costa, A. et al. (2020). Ensemble-based data assimilation of volcanic ash clouds from satellite observations: Application to the 24 December 2018 Mt. Etna Explosive Eruption. Atmosphere, 11(4), 359.Google Scholar
Pavolonis, M. J., Feltz, W. F., Heidinger, A. K., and Gallina, G. M. (2006). A daytime complement to the reverse absorption technique for improved automated detection of volcanic ash. Journal of Atmospheric and Oceanic Technology, 23, 1422–44.Google Scholar
Pelley, R., Cooke, M., Manning, A. et al. (2015). Initial implementation of an inversion technique for estimating volcanic ash source parameters in near real time using satellite retrievals: Forecasting Research Technical Report, vol. 604. Met Exeter, UK: Met Office. https://library.metoffice.gov.uk/Portal/Default/en-GB/DownloadImageFile.ashx?objectId=415&ownerType=0&ownerId=212804.Google Scholar
Prata, A. J. (1989). Observations of volcanic ash clouds in the 10–12-micron window using AVHRR/2 Data. International Journal of Remote Sensing, 10, 751–61.Google Scholar
Prata, A. T. (2016). Remote sensing of volcanic eruptions. In Duarte, J. C. and Schellart, W. P., eds., Plate Boundaries and Natural Hazards, American Geophysical Union (AGU). Hoboken, NJ: Wiley and Sons, pp. 289322.Google Scholar
Prata, F., and Lynch, M. (2019). Passive Earth observations of volcanic clouds in the atmosphere. Atmosphere, 10, 199.Google Scholar
Prata, A. T., Mingari, L., Folch, A., Macedonio, G., and Costa, A. (2021). FALL3D-8.0: A computational model for atmospheric transport and deposition of particles, aerosols and radionuclides. Part 2: Model validation. Geoscientific Model Development, 14, 409–36.Google Scholar
Robock, A., and Oppenheimer, C. (2003). Volcanism and the Earth’s atmosphere: American Geophysical Union, Geophysical Monograph, 139, 360 pp.Google Scholar
Seibert, P. (2000). Inverse modelling of sulfur emissions in Europe based on trajectories. In Kasibhatla, P., Heimann, M., Rayner, P., Mahowald, N., Prinn, R. G., and Hartley, D. E, eds., Inverse Methods in Global Biogeochemical Cycles, Geophysical Monograph 114. Washington, DC: American Geophysical Union, pp. 147–54,Google Scholar
Sparks, R. S. J., Bursik, M. I., Carey, S. N. et al. (1997). Volcanic Plumes. Chichester: John Wiley & Sons.Google Scholar
Steensen, B. M., Kylling, A., Kristiansen, N. I., and Schulz, M. (2017). Uncertainty assessment and applicability of an inversion method for volcanic ash forecasting. Atmospheric Chemistry and Physics, 17, 9205–22.Google Scholar
Stohl, A., Prata, A., Eckhardt, S. et al. (2011). Determination of time- and height-resolved volcanic ash emissions and their use for quantitative ash dispersion modeling: The 2010 Eyjafjallajokull eruption. Atmospheric Chemistry and Physics, 11, 4333–51.Google Scholar
Suzuki, T. (1983). A theoretical model for dispersion of tephra. In Shimozuru, D. and Yokoyama, I, eds., Volcanism: Physics and Tectonics. Tokyo: Arc, pp. 95113.Google Scholar
Theys, N., Hedelt, P., De Smedt, I. et al. (2019). Global monitoring of volcanic SO2 degassing with unprecedented resolution from TROPOMI onboard Sentinel-5 Precursor. Scientific Reports, 9, 2643.Google Scholar
Van Leeuwen, P. J., and Ades, M. (2013). Efficient fully nonlinear data assimilation for geophysical fluid dynamics. Computers & Geosciences, 55, 1627.Google Scholar
Vira, J., Carboni, E., Grainger, R. G., and Sofiev, M. (2017). Variational assimilation of IASI SO2 plume height and total column retrievals in the 2010 eruption of Eyjafjallajökull using the SILAM v5.3 chemistry transport model.Geoscientific Model Development, 10, 19852008.Google Scholar
Wilkins, K. L., Mackie, S., Watson, M. et al. (2014). Data insertion in volcanic ash cloud forecasting. Annals of Geophysics, Fast Track, 2. https://doi.org/10.4401/ag-6624.Google Scholar
Wilkins, K. L., Watson, I. M., Kristiansen, N. I. et al. (2016). Using data insertion with the NAME model to simulate the 8 May 2010 Eyjafjallajökull volcanic ash cloud. Journal of Geophysical Research: Atmospheres, 121, 306–23. https://doi.org/10.1002/2015JD023895.Google Scholar
Zhou, H., Gómez-Hernández, J., Hendricks Franssen, H. J., and Li, L. (2011). An approach to handling non-Gaussianity of parameters and state variables in ensemble Kalman filtering. Advances in Water Resources, 34(7), 844–64.CrossRefGoogle Scholar
Zidikheri, M. J., and Potts, R. J. (2015). A simple inversion method for determining optimal dispersion model parameters from satellite detections of volcanic sulfur dioxide. Journal of Geophysical Research: Atmospheres, 120, 9702–17.Google Scholar
Zidikheri, M. J., and Lucas, C. (2021a). A computationally efficient ensemble filtering scheme for quantitative volcanic ash forecasts. Journal of Geophysical Research: Atmospheres, 126, e2020JD033094.Google Scholar
Zidikheri, M. J., and Lucas, C. (2021b). Improving ensemble volcanic ash forecasts by direct insertion of satellite data and ensemble filtering. Atmosphere, 12, 1215.CrossRefGoogle Scholar
Zidikheri, M., Potts, R. J., and Lucas, C. (2016). A probabilistic inverse method for volcanic ash dispersion modelling. The ANZIAM Journal, 56, 194209.Google Scholar
Zidikheri, M., Lucas, C., and Potts, R. (2017a), Estimation of optimal dispersion model source parameters using satellite detections of volcanic ash. Journal of Geophysical Research: Atmospheres, 122, 8207–32.Google Scholar
Zidikheri, M., Lucas, C., and Potts, R. (2017b). Toward quantitative forecasts of volcanic ash dispersal: Using satellite retrievals for optimal estimation of source terms. Journal of Geophysical Research: Atmospheres, 122, 8187–206.Google Scholar

References

Albert, J. (2005). Evaluation of quasi-linear diffusion coefficients for whistler mode waves in a plasma with arbitrary density ratio. Journal of Geophysical Research: Space Physics, 110(A3).Google Scholar
Aseev, N., and Shprits, Y. (2019). Reanalysis of ring current electron phase space densities using Van Allen Probe observations, convection model, and log-normal Kalman Filter. Space Weather, 17(4), 619–38.Google Scholar
Aseev, N. A., Shprits, Y. Y., Drozdov, A. Y., and Kellerman, A. C. (2016). Numerical applications of the advective-diffusive codes for the inner magnetosphere. Space Weather, 14(11), 9931010.Google Scholar
Asikainen, T., and Mursula, K. (2011). Recalibration of the long-term NOAA/MEPED energetic proton measurements. Journal of Atmospheric and Solar-Terrestrial Physics, 73(2), 335–47.Google Scholar
Asikainen, T., and Mursula, K. (2013). Correcting the NOAA/MEPED energetic electron fluxes for detector efficiency and proton contamination. Journal of Geophysical Research: Space Physics, 118(10), 6500–10.Google Scholar
Baker, D., Belian, R., Higbie, P., Klebesadel, R., and Blake, J. (1987). Deep dielectric charging effects due to high-energy electrons in Earth’s outer magnetosphere. Journal of Electrostatics, 20(1), 319.CrossRefGoogle Scholar
Baker, D. N. (2000). The occurrence of operational anomalies in spacecraft and their relationship to space weather. IEEE Transactions on Plasma Science, 28(6), 2007–16.Google Scholar
Baker, D. N. (2002). How to cope with space weather. Science, 297(5586), 1486–7.CrossRefGoogle ScholarPubMed
Baker, D. N. (2005). Specifying and forecasting space weather threats to human technology. In Daglis, I. A., ed., Effects of Space Weather on Technology Infrastructure, NATO Science Series II: Mathematics, Physics and Chemistry, vol. 176. Dordrecht: Springer, pp. 125.Google Scholar
Beutier, T., and Boscher, D. (1995). A three-dimensional analysis of the electron radiation belt by the Salammbô code. Journal of Geophysical Research: Space Physics, 100(A8), 14853–61.Google Scholar
Bourdarie, S., and Maget, V. (2012). Electron radiation belt data assimilation with an ensemble Kalman filter relying on the Salammbô code. Annales Geophysicae, 30, 929–43.Google Scholar
Castillo, A. M., de Wiljes, J., Shprits, Y. Y., and Aseev, N. A. (2021). Reconstructing the dynamics of the outer electron radiation belt by means of the standard and ensemble Kalman filter with the VERB-3D code. Space Weather, 19(10), e2020SW002672.Google Scholar
Cervantes, S., Shprits, Y. Y., Aseev, N. et al. (2020a). Identifying radiation belt electron source and loss processes by assimilating spacecraft data in a three-dimensional diffusion model. Journal of Geophysical Research: Space Physics, 125(1), 116.Google Scholar
Cervantes, S., Shprits, Y. Y., Aseev, N. A., and Allison, H. J. (2020b). Quantifying the effects of EMIC wave scattering and magnetopause shadowing in the outer electron radiation belt by means of data assimilation. Journal of Geophysical Research: Space Physics, 125(8).Google Scholar
Cohen, D., Spanjers, G., Winter, J. et al. (2005). Design and Systems Engineering of AFRL’s Demonstration and Sciences Experiment. Proceedings of the GATech Space Systems Engineering Conference, 8–10 November 2005. Paper No. GT-SSEC.D.1, pp. 17. http://hdl.handle.net/1853/8037.Google Scholar
Daae, M., Shprits, Y. Y., Ni, B. et al. (2011). Reanalysis of radiation belt electron phase space density using various boundary conditions and loss models. Advances in Space Research, 48(8), 1327–34.Google Scholar
Drozdov, A. Y., Shprits, Y. Y., Orlova, K. G. et al. (2015). Energetic, relativistic, and ultrarelativistic electrons: Comparison of long-term VERB code simulations with Van Allen Probes measurements. Journal of Geophysical Research: Space Physics, 120(5), 3574–87.Google Scholar
Escoubet, C., Schmidt, R., and Goldstein, M. eds. (1997). Cluster-science and mission overview. In Escoubet, C. P., Russell, C. T., and Schmidt, R., eds., The Cluster and Phoenix Missions. Dordrecht: Springer, pp. 1132.Google Scholar
Evensen, G. (2003). The ensemble Kalman filter: Theoretical formulation and practical implementation. Ocean Dynamics, 53(4), 343–67.Google Scholar
Fok, M.-C., Glocer, A., Zheng, Q. et al. (2011). Recent developments in the radiation belt environment model. Journal of Atmospheric and Solar-Terrestrial Physics, 73(11–12), 1435–43.Google Scholar
Fok, M.-C., Horne, R. B., Meredith, N. P., and Glauert, S. A. (2008). Radiation belt environment model: Application to space weather nowcasting. Journal of Geophysical Research: Space Physics, 113(A3).Google Scholar
Friedel, R., Bourdarie, S., Fennell, J., Kanekal, S., and Cayton, T. (2003). ‘Nudging’ the Salammbo Code: First results of seeding a diffusive radiation belt code with in situ data: GPS, GEO, HEO and POLAR. Eos Transactions of the American Geophysical Union, 84(46), Fall Meeting, 8–12 December 2003, San Francisco, 476 Suppl., abstract, SM11D–06.Google Scholar
Friedel, R., Reeves, G., and Obara, T. (2002). Relativistic electron dynamics in the inner magnetosphere: A review. Journal of Atmospheric and Solar-Terrestrial Physics, 64, 265–82.Google Scholar
Galand, M., and Evans, D. S. (2000). Radiation damage of the proton MEPED detector on POES (TIROS/NOAA) satellites. NOAA Technical Memorandum. Boulder, CO. OAR 456-SEC 42.Google Scholar
Ganushkina, N. Y., Amariutei, O., Shprits, Y., and Liemohn, M. (2013). Transport of the plasma sheet electrons to the geostationary distances. Journal of Geophysical Research: Space Physics, 118(1), 8298.Google Scholar
Ganushkina, N. Y., Liemohn, M., Amariutei, O., and Pitchford, D. (2014). Low-energy electrons (5–50 kev) in the inner magnetosphere. Journal of Geophysical Research: Space Physics, 119(1), 246–59.Google Scholar
Glauert, S. A., and Horne, R. B. (2005). Calculation of pitch angle and energy diffusion coefficients with the PADIE code. Journal of Geophysical Research: Space Physics, 110(A4).Google Scholar
Godinez, H. C., Yu, Y., Lawrence, E. et al. (2016). Ring current pressure estimation with RAM-SCB using data assimilation and Van Allen Probe flux data. Geophysical Research Letters, 43(23), 11948–56.Google Scholar
Green, J. C., and Kivelson, M. G. (2004). Relativistic electrons in the outer radiation belt: Differentiating between acceleration mechanisms. Journal of Geophysical Research: Space Physics, 109(A3).Google Scholar
Hamlin, D. A., Karplus, R., Vik, R. C., and Watson, K. M. (1961). Mirror and azimuthal drift frequencies for geomagnetically trapped particles. Journal of Geophysical Research (1896–1977), 66(1), 14.Google Scholar
Jordanova, V., Kozyra, J., Nagy, A., and Khazanov, G. (1997). Kinetic model of the ring current–atmosphere interactions. Journal of Geophysical Research: Space Physics, 102(A7), 14279–91.CrossRefGoogle Scholar
Jordanova, V., and Miyoshi, Y. (2005). Relativistic model of ring current and radiation belt ions and electrons: Initial results. Geophysical Research Letters, 32(14).Google Scholar
Kalman, R. (1960). A new approach to linear filtering and prediction problems. Trans. ASME Journal of Basic Engineering, 82(1), 3545.Google Scholar
Kalnay, E. (2003). Atmospheric Modeling, Data Assimilation and Predictability. Cambridge: Cambridge University Press.Google Scholar
Kennel, C., and Engelmann, F. (1966). Velocity space diffusion from weak plasma turbulence in a magnetic field. The Physics of Fluids, 9(12), 2377–88.Google Scholar
Kim, K., Shprits, Y., Subbotin, D., and Ni, B. (2012). Relativistic radiation belt electron responses to GEM magnetic storms: Comparison of CRRES observations with 3-D VERB simulations. Journal of Geophysical Research: Space Physics, 117(A8).Google Scholar
Koller, J., Chen, Y., Reeves, G. D. et al. (2007). Identifying the radiation belt source region by data assimilation. Journal of Geophysical Research: Space Physics, 112(A6).Google Scholar
Koller, J., Friedel, R., and Reeves, G. (2005). Radiation belt data assimilation and parameter estimation. LANL Reports, LA-UR-05-6700. http://library.lanl.gov/cgi-bin/getfile?LA-UR-05-6700.pdf.Google Scholar
Kondrashov, D., Ghil, M., and Shprits, Y. Y. (2011). Lognormal Kalman filter for assimilating phase space density data in the radiation belts. Space Weather, 9(11).Google Scholar
Kondrashov, D., Shprits, Y. Y., Ghil, M., and Thorne, R. (2007). A Kalman filter technique to estimate relativistic electron lifetimes in the outer radiation belt. Journal of Geophysical Research: Space Physics, 112(A10).Google Scholar
Lanzerotti, L. J. (2001). Space weather effects on technologies. Washington DC American Geophysical Union Geophysical Monograph Series, 125, 1122.Google Scholar
Lyons, L. R., Thorne, R. M., and Kennel, C. F. (1972). Pitch-angle diffusion of radiation belt electrons within the plasmasphere. Journal of Geophysical Research, 77(19), 3455–74.Google Scholar
Maynard, N. C., and Chen, A. J. (1975). Isolated cold plasma regions: Observations and their relation to possible production mechanisms. Journal of Geophysical Research (1896–1977), 80(7), 1009–13.Google Scholar
McFadden, J., Evans, D., Kasprzak, W. et al. (2007). In-flight instrument calibration and performance verification. In Wüest, M., Evansvon, D. S. and Steiger, R., eds., Calibration of Particle Instruments in Space Physics, vol. SR-007. Bern: International Space Science Institute, pp. 277385.Google Scholar
Miyoshi, Y., Shinohara, I., Takashima, T. et al. (2018). Geospace exploration project ERG. Earth, Planets and Space, 70(1), 113.Google Scholar
Naehr, S. M., and Toffoletto, F. R. (2005). Radiation belt data assimilation with an extended Kalman filter. Space Weather, 3(6).Google Scholar
Nakano, S., Ueno, G., Ebihara, Y. et al. (2008). A method for estimating the ring current structure and the electric potential distribution using energetic neutral atom data assimilation. Journal of Geophysical Research: Space Physics, 113(A5).Google Scholar
Ni, B., Shprits, Y., Nagai, T. et al. (2009a). Reanalyses of the radiation belt electron phase space density using nearly equatorial CRRES and polar-orbiting Akebono satellite observations. Journal of Geophysical Research: Space Physics, 114(A5).Google Scholar
Ni, B., Shprits, Y., Thorne, R., Friedel, R., and Nagai, T. (2009b). Reanalysis of relativistic radiation belt electron phase space density using multisatellite observations: Sensitivity to empirical magnetic field models. Journal of Geophysical Research: Space Physics, 114(A12).Google Scholar
Ni, B., Shprits, Y. Y., Friedel, R. H. et al. (2013). Responses of Earth’s radiation belts to solar wind dynamic pressure variations in 2002 analyzed using multisatellite data and Kalman filtering. Journal of Geophysical Research: Space Physics, 118(7), 4400–14.Google Scholar
Podladchikova, T. V., Shprits, Y. Y., Kellerman, A. C., and Kondrashov, D. (2014a). Noise statistics identification for Kalman filtering of the electron radiation belt observations: 2. Filtration and smoothing. Journal of Geophysical Research: Space Physics, 119(7), 5725–43.Google Scholar
Podladchikova, T. V., Shprits, Y. Y., Kondrashov, D., and Kellerman, A. C. (2014b). Noise statistics identification for Kalman filtering of the electron radiation belt observations: 1. Model errors. Journal of Geophysical Research: Space Physics, 119(7), 5700–24.Google Scholar
Reeves, G. D., McAdams, K. L., Friedel, R. H. W., and O’Brien, T. P. (2003). Acceleration and loss of relativistic electrons during geomagnetic storms. Geophysical Research Letters, 30(10).Google Scholar
RobinsonJr., P. A. (1989). Spacecraft environmental anomalies handbook. Technical report, Jet Propulsion Lab, Pasadena, CA.Google Scholar
Roederer, J. G. (2012). Dynamics of Geomagnetically Trapped radiation, vol. 2. Berlin: Springer Science & Business Media.Google Scholar
Rosen, A. (1976). Spacecraft charging by magnetospheric plasmas. IEEE Transactions on Nuclear Science, 23(6), 1762–68.Google Scholar
Saikin, A. A., Shprits, Y. Y., Drozdov, A. Y. et al. (2021). Reconstruction of the radiation belts for solar cycles 17–24 (1933–2017). Space Weather, 19(3), e2020SW002524.Google Scholar
Schiller, Q., Li, X., Koller, J., Godinez, H., and Turner, D. L. (2012). A parametric study of the source rate for outer radiation belt electrons using a Kalman filter. Journal of Geophysical Research: Space Physics, 117(A9).Google Scholar
Schulz, M., and Lanzerotti, L. (1974). Particle Diffusion in the Radiation Belts. New York: Springer-Verlag.Google Scholar
Shprits, Y., Daae, M., and Ni, B. (2012). Statistical analysis of phase space density buildups and dropouts. Journal of Geophysical Research: Space Physics, 117(A1).Google Scholar
Shprits, Y., Elkington, S., Meredith, N., and Subbotin, D. (2008a). Review of modeling of losses and sources of relativistic electrons in the outer radiation belts: I. Radial transport. Journal of Atmospheric and Solar-Terrestrial Physics, 70(14), 1679–93.Google Scholar
Shprits, Y. Y., Kellerman, A. C., Drozdov, A. Y. et al. (2015). Combined convective and diffusive simulations: VERB-4D comparison with 17 March 2013 Van Allen Probes observations. Geophysical Research Letters, 42(22), 9600–8.Google Scholar
Shprits, Y., Kellerman, A., Kondrashov, D., and Subbotin, D. (2013). Application of a new data operator-splitting data assimilation technique to the 3-D VERB diffusion code and CRRES measurements. Geophysical Research Letters, 40(19), 49985002.Google Scholar
Shprits, Y., Kondrashov, D., Chen, Y. et al. (2007). Reanalysis of relativistic radiation belt electron fluxes using CRRES satellite data, a radial diffusion model, and a Kalman filter. Journal of Geophysical Research: Space Physics, 112(A12216).Google Scholar
Shprits, Y., Subbotin, D., Meredith, N., and Elkington, S. (2008b). Review of modeling of losses and sources of relativistic electrons in the outer radiation belts: II. Local acceleration and loss. Journal of Atmospheric and Solar-Terrestrial Physics, 70(14), 1694–713.Google Scholar
Shprits, Y., Subbotin, D., and Ni, B. (2009). Evolution of electron fluxes in the outer radiation belt computed with the VERB code Journal of Geophysical Research: Space Physics, 114(A11).Google Scholar
Shprits, Y., Subbotin, D., Ni, B. et al. (2011). Profound change of the near-Earth radiation environment caused by solar superstorms. Space Weather, 9(8).Google Scholar
Shprits, Y., Thorne, R., Friedel, R. et al. (2006). Outward radial diffusion driven by losses at magnetopause. Journal of Geophysical Research: Space Physics, 111(A11).Google Scholar
Shprits, Y. Y., Vasile, R., and Zhelavskaya, I. S. (2019). Nowcasting and predicting the Kp index using historical values and real-time observations. Space Weather, 17(8), 1219–29.Google Scholar
Sibeck, D., and Angelopoulos, V. (2008). THEMIS science objectives and mission phases. Space Science Reviews, 141(1), 3559.Google Scholar
Stern, D. P. (1975). The motion of a proton in the equatorial magnetosphere. Journal of Geophysical Research (1896–1977), 80(4), 595–9.Google Scholar
Stix, T. H. (1962). The Theory of Plasma Waves. New York: McGraw-Hill.Google Scholar
Subbotin, D. A. and Shprits, Y. Y. (2009). Three-dimensional modeling of the radiation belts using the Versatile Electron Radiation Belt (VERB) code. Space Weather, 7(10).Google Scholar
Subbotin, D. A. and Shprits, Y. Y. (2012). Three-dimensional radiation belt simulations in terms of adiabatic invariants using a single numerical grid. Journal of Geophysical Research: Space Physics, 117(A05205).Google Scholar
Subbotin, D., Shprits, Y., and Ni, B. (2011). Long-term radiation belt simulation with the VERB 3-D code: Comparison with CRRES observations. Journal of Geophysical Research: Space Physics, 116(A12).Google Scholar
Tsyganenko, N. (1989). A magnetospheric magnetic field model with a warped tail current sheet. Planetary and Space Science, 37(1), 5–20.Google Scholar
Turner, D., Shprits, Y., Hartinger, M., and Angelopoulos, V. (2012). Explaining sudden losses of outer radiation belt electrons during geomagnetic storms. Nature Physics, 8(3), 208–12.Google Scholar
UCS (2021). UCS Satellite Database, Union of Concerned Scientists. https://ucsusa.org/resources/satellite-database/.Google Scholar
Volland, H. (1973). A semiempirical model of large-scale magnetospheric electric fields. Journal of Geophysical Research (1896–1977), 78(1), 171–80.Google Scholar
Wang, D. and Shprits, Y. Y. (2019). On how high-latitude chorus waves tip the balance between acceleration and loss of relativistic electrons. Geophysical Research Letters, 46(14), 7945–54.Google Scholar
Wang, D., Shprits, Y. Y., Zhelavskaya, I. S. et al. (2020). The effect of plasma boundaries on the dynamic evolution of relativistic radiation belt electrons. Journal of Geophysical Research: Space Physics, 125(5), e2019JA027422.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×