More stories

  • in

    Reevaluating streamflow declines across the middle East and central Asia with insights from change point detection

    Abstract

    Change point detection (CPD) methods are widely used to identify abrupt shifts in streamflow, often linked to human activities. This study assesses the effectiveness of CPD techniques across countries in the Middle East, Central Asia, and Pakistan—regions vulnerable to water scarcity and climate variability. Analysis of annual streamflow data (1970–2018) revealed that 1998 was the most frequent breakpoint, coinciding with a major El Niño event, suggesting climatic anomalies were key drivers of the initial change. The coherence of breakpoints across the region highlights the influence of large-scale climate signals. However, comparison of pre- and post-break conditions indicates that the magnitude of hydrological shifts cannot be explained by climate alone, highlighting the limitations of CPD methods in distinguishing climatic from anthropogenic drivers. To explore these dynamics in more detail, the Karkheh River Basin (KRB) in Iran was examined. A marked change in streamflow patterns around 1998 was observed, along with shifts in temperature and precipitation. These results underscore the need for cautious use of CPD methods when attributing hydrological changes to human versus climatic factors.

    Data availability

    The data used in this study is owned by the Iran Water Resources Management Organization (https://www.wrm.ir), where the datasets can be accessed.
    ReferencesGreen, T. R., Taniguchi, M. & Kooi, H. Potential impacts of climate change and human activity on subsurface water resources. Vadose Zone J. 6, 531–532 (2007).
    Google Scholar 
    Ryberg, K. R., Hodgkins, G. A. & Dudley, R. W. Change points in annual peak streamflows: method comparisons and historical change points in the united States. J. Hydrol. (Amst). 583, 124307 (2020).
    Google Scholar 
    Dey, P. & Mishra, A. Separating the impacts of climate change and human activities on streamflow: A review of methodologies and critical assumptions. J. Hydrol. (Amst). 548, 278–290 (2017).
    Google Scholar 
    Masson-Delmotte, V. et al. Climate change 2021: the physical science basis. Contribution Working group. I Sixth Assess. Rep. Intergovernmental Panel Clim. Change. 2, 2391 (2021).
    Google Scholar 
    Huang, J. C., Lee, T. Y. & Lee, J. Y. Observed magnified runoff response to rainfall intensification under global warming. Environ. Res. Lett. 9, 034008 (2014).
    Google Scholar 
    Wrzesien, M. L. & Pavelsky, T. M. Projected changes to extreme runoff and precipitation events from a downscaled simulation over the Western united States. Front. Earth Sci. (Lausanne) 7, 355 (2020).
    Google Scholar 
    Wen, S. et al. Attribution of streamflow changes during 1961–2019 in the upper Yangtze and the upper yellow river basins. Clim. Change. 177, 60 (2024).
    Google Scholar 
    Liu, J., Zhang, C., Kou, L. & Zhou, Q. Effects of climate and land use changes on water resources in the Taoer river. Adv. Meteorol. 2017, 1–13 (2017).
    Google Scholar 
    Tan, X., Liu, B., Tan, X., Huang, Z. & Fu, J. Combined impacts of climate change and human activities on blue and green water resources in a high-intensity development watershed. Hydrol. Earth Syst. Sci. 29, 427–445 (2025).
    Google Scholar 
    Zhang, W., Villarini, G., Vecchi, G. A. & Smith, J. A. Urbanization exacerbated the rainfall and flooding caused by hurricane Harvey in Houston. Nature 563, 384–388 (2018).
    Google Scholar 
    Zeng, F. et al. Reduced runoff in the upper Yangtze river due to comparable contribution of anthropogenic and climate changes. Earths Future 12, e2023EF004028 (2024).
    Google Scholar 
    Li, Z., Li, Q., Wang, J., Feng, Y. & Shao, Q. Impacts of projected climate change on runoff in upper reach of Heihe river basin using climate elasticity method and GCMs. Sci. Total Environ. 716, 137072 (2020).
    Google Scholar 
    Shah, L. A. et al. Statistical significance assessment of streamflow elasticity of major rivers. Civil Eng. J. 7, 893–905 (2021).
    Google Scholar 
    Somorowska, U. & Łaszewski, M. Quantifying streamflow response to climate variability, wastewater inflow, and sprawling urbanization in a heavily modified river basin. Sci. Total Environ. 656, 458–467 (2019).
    Google Scholar 
    Yu, K. et al. Evaluating the impact of ecological construction measures on water balance in the loess plateau region of China within the Budyko framework. J. Hydrol. (Amst). 601, 126596 (2021).
    Google Scholar 
    Friedman, D. et al. US army corps of engineers nonstationarity detection tool user guide (US Army Corps of Engineers: Washington, DC 2016). Koutsoyiannis, D. Nonstationarity versus scaling in hydrology. J. Hydrol. (Amst). 324, 239–254 (2006).
    Google Scholar 
    Stosic, T., Telesca, L., de Souza Ferreira, D. V. & Stosic, B. Investigating anthropically induced effects in streamflow dynamics by using permutation entropy and statistical complexity analysis: A case study. J. Hydrol. (Amst). 540, 1136–1145 (2016).
    Google Scholar 
    Zhang, Q., Singh, V. P., Li, K. & Li, J. Trend, periodicity and abrupt change in streamflow of the East river, the Pearl river basin. Hydrol. Process. 28, 305–314 (2014).
    Google Scholar 
    Buishand, T. A. Some methods for testing the homogeneity of rainfall records. J. Hydrol. (Amst). 58, 11–27 (1982).
    Google Scholar 
    Alexandersson, H. & institutionen, U. M. A Homogeneity Test Based on Ratios and Applied To Precipitation Series (Inst., Univ., 1984). Pettitt, A. N. A Non-Parametric approach to the Change-Point problem. Appl. Stat. 28, 126 (1979).
    Google Scholar 
    Animashaun, I. M., Oguntunde, P. G., Akinwumiju, A. S. & Olubanjo, O. O. Rainfall analysis over the Niger central hydrological Area, Nigeria: Variability, trend, and change point detection. Sci. Afr. 8, e00419 (2020).
    Google Scholar 
    Dariane, A. B. & Pouryafar, E. Quantifying and projection of the relative impacts of climate change and direct human activities on streamflow fluctuations. Clim. Change. 165, 34 (2021).
    Google Scholar 
    Hou, M. et al. Streamflow composition and the contradicting impacts of anthropogenic activities and Climatic change on streamflow in the Amu Darya basin, central Asia. J. Hydrometeorol. 24, 185–201 (2023).
    Google Scholar 
    Jiang, C., Zhang, L. & Tang, Z. Multi-temporal scale changes of streamflow and sediment discharge in the headwaters of yellow river and Yangtze river on the Tibetan Plateau, China. Ecol. Eng. 102, 240–254 (2017).
    Google Scholar 
    Li, M., Gu, H., Wang, H., Wang, Y. & Chi, B. Quantifying the impact of climate variability and human activities on streamflow variation in Taoer river Basin, China. Environ. Sci. Pollut. Res. 30, 56425–56439 (2023).
    Google Scholar 
    Minaei, M. & Irannezhad, M. Spatio-temporal trend analysis of precipitation, temperature, and river discharge in the Northeast of Iran in recent decades. Theor. Appl. Climatol. 131, 167–179 (2018).
    Google Scholar 
    Xie, P., Gu, H., Sang, Y. F., Wu, Z. & Singh, V. P. Comparison of different methods for detecting change points in hydroclimatic time series. J. Hydrol. (Amst). 577, 123973 (2019).
    Google Scholar 
    Zuo, D., Xu, Z., Wu, W., Zhao, J. & Zhao, F. Identification of streamflow response to climate change and human activities in the Wei river Basin, China. Water Resour. Manag. 28, 833–851 (2014).
    Google Scholar 
    Perreault, L., Bernier, J., Bobée, B. & Parent, E. Bayesian change-point analysis in hydrometeorological time series. Part 1. The normal model revisited. J. Hydrol. (Amst) 235, 221–241 (2000).
    Google Scholar 
    Perreault, L., Bernier, J., Bobée, B. & Parent, E. Bayesian change-point analysis in hydrometeorological time series. Part 2. Comparison of change-point models and forecasting. J. Hydrol. (Amst) 235, 242–263 (2000).
    Google Scholar 
    Yang, L. et al. Runoff changes in the major river basins of China and their responses to potential driving forces. J. Hydrol. (Amst). 607, 127536 (2022).
    Google Scholar 
    Zeng, F., Ma, M. G., Di, D. R. & Shi, W. Y. Separating the impacts of climate change and human activities on runoff: A review of method and application. Water (Basel). 12, 2201 (2020).
    Google Scholar 
    Nourani, V., Gholinia, A., Abbaszadeh, P., Ahmadi, R. & Ke, C. Q. Unravelling the impact of climate change and anthropogenic activities on streamflow: The benefit of newly developed evapotranspiration data. Hydrol. Sci. J. 69, 1–18 (2024). https://doi.org/10.1080/02626667.2024.2398654 (2024).
    Google Scholar 
    Al-Hasani, A. A. J. Trend analysis and abrupt change detection of streamflow variations in the lower Tigris river Basin, Iraq. Int. J. River Basin Manag. 19, 523–534 (2021).
    Google Scholar 
    Bai, X. & Zhao, W. Impacts of climate change and anthropogenic stressors on runoff variations in major river basins in China since 1950. Sci. Total Environ. 898, 165349 (2023).
    Google Scholar 
    Banda, V. D., Dzwairo, R. B., Singh, S. K. & Kanyerere, T. Separating anthropogenic and climate contributions to streamflow variations in Rietspruit sub-basin, South Africa. Phys. Chem. Earth Parts A/B/C. 127, 103200 (2022).
    Google Scholar 
    Darvini, G. & Memmola, F. Assessment of the impact of climate variability and human activities on the runoff in five catchments of the Adriatic Coast of south-central Italy. J. Hydrol. Reg. Stud. 31, 100712 (2020).
    Google Scholar 
    Hu, Y. et al. An integrated assessment of runoff dynamics in the Amu Darya river basin: confronting climate change and multiple human activities, 1960–2017. J. Hydrol. (Amst) 603, 126905 (2021).
    Google Scholar 
    Jiang, C. & Wang, F. Temporal changes of streamflow and its causes in the Liao river basin over the period of 1953–2011, Northeastern China. Catena (Amst). 145, 227–238 (2016).
    Google Scholar 
    Liu, N., Harper, R. J., Smettem, K. R. J., Dell, B. & Liu, S. Responses of streamflow to vegetation and climate change in Southwestern Australia. J. Hydrol. (Amst). 572, 761–770 (2019).
    Google Scholar 
    Panditharathne, R. et al. Trends and variabilities in rainfall and streamflow: A case study of the Nilwala river basin in Sri Lanka. Hydrology 10, 8 (2022).
    Google Scholar 
    Saifullah, M. et al. Hydrological response of the Kunhar river basin in Pakistan to climate change and anthropogenic impacts on runoff characteristics. Water (Basel). 13, 3163 (2021).
    Google Scholar 
    Wu, L. et al. Evaluating the contributions of climate change and human activities to runoff in typical semi-arid area, China. J. Hydrol. (Amst). 590, 125555 (2020).
    Google Scholar 
    Zhang, Y. et al. Analysis of impacts of climate variability and human activity on streamflow for a river basin in Northeast China. J. Hydrol. (Amst). 410, 239–247 (2011).
    Google Scholar 
    Thielen, D. R. et al. Effect of extreme El Niño events on the precipitation of Ecuador. Nat. Hazards Earth Syst. Sci. 23, 1507–1527 (2023).
    Google Scholar 
    Milly, P. C. D. et al. On critiques of stationarity is dead: whither water management? Water Resour. Res. 51, 7785–7789 (2015).
    Google Scholar 
    Sen, P. K. Estimates of the regression coefficient based on kendall’s Tau. J. Am. Stat. Assoc. 63, 1379–1389 (1968).
    Google Scholar 
    Wang, G., Xia, J. & Chen, J. Quantification of effects of climate variations and human activities on runoff by a monthly water balance model: A case study of the Chaobai river basin in Northern China. Water Resour. Res 45 (2009). Kazemzadeh, M. & Malekian, A. Changeability evaluation of hydro-climate variables in Western Caspian sea region, Iran. Environ. Earth Sci. 77, 120 (2018).
    Google Scholar 
    Sharifi, A., Mirabbasi, R., Nasr-Esfahani, A. & Torabi Haghighi, M. Fatahi Nafchi, R. Quantifying the impacts of anthropogenic changes and climate variability on runoff changes in central plateau of Iran using nine methods. J. Hydrol. (Amst). 603, 127045 (2021).
    Google Scholar 
    Mikaeili, O. & Shourian, M. Assessment of the analytic and hydrologic methods in separation of watershed response to climate and land use changes. Water Resour. Manage. 37, 2575–2591 (2023).
    Google Scholar 
    Jalali, J., Ahmadi, A. & Abbaspour, K. Runoff responses to human activities and climate change in an arid watershed of central Iran. Hydrol. Sci. J. 66, 2280–2297 (2021).
    Google Scholar 
    Ansari Mahabadi, S. & Delavar, M. Evaluation and comparison of different methods for determining the contribution of Climatic factors and direct human interventions in reducing watershed discharge. Ecol. Indic. 158, 111480 (2024).
    Google Scholar 
    Latif, M., Semenov, V. A. & Park, W. Super El Niños in response to global warming in a climate model. Clim. Change. 132, 489–500 (2015).
    Google Scholar 
    Varotsos, C. A. & Deligiorgi, D. G. Sea-surface temperature and Southern Oscillation signal in the upper stratosphere‐lower mesosphere. Int. J. Climatol. 11, 77–83 (1991).
    Google Scholar 
    Alizadeh-Choobari, O. & Najafi, M. S. Extreme weather events in Iran under a changing climate. Clim. Dyn. 50, 249–260 (2018).
    Google Scholar 
    Barlow, M. et al. A review of drought in the middle East and Southwest Asia. J. Clim. 29, 8547–8574 (2016).
    Google Scholar 
    Salehi, S., Dehghani, M., Mortazavi, S. M. & Singh, V. P. Trend analysis and change point detection of seasonal and annual precipitation in Iran. Int. J. Climatol. 40, 308–323 (2020).
    Google Scholar 
    Zarezadeh, M., Delavar, M., Morid, S. & Abbasi, H. Evaluating the effectiveness of macro-level water-saving policies based on water footprint sustainability indicators. Agric. Water Manag. 282, 108272 (2023).
    Google Scholar 
    Zarenistanak, M., Kripalani, R. H. & Dhorde, A. G. Trend analysis and change point detection of annual and seasonal precipitation and temperature series over Southwest Iran. J. Earth Syst. Sci. 123, 281–295 (2014).
    Google Scholar 
    Download referencesAuthor informationAuthors and AffiliationsDepartment of Civil Engineering, K.N.Toosi University of Technology, Tehran, IranAlireza Borhani Dariane & Mahboobeh GhasemiAuthorsAlireza Borhani DarianeView author publicationsSearch author on:PubMed Google ScholarMahboobeh GhasemiView author publicationsSearch author on:PubMed Google ScholarContributionsA.B.D. supervised the research, developed the main idea and methodology, and analyzed the results. M.Gh. prepared the computer models, obtained the results and wrote the initial draft, which was then modified and finalized by A.B.D.Corresponding authorCorrespondence to
    Alireza Borhani Dariane.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Ethical approval
    All authors have read, understood, and have complied as applicable with the statement on “Ethical responsibilities of Authors” as found in the Instructions for Authors.

    Additional informationPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Rights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleDariane, A.B., Ghasemi, M. Reevaluating streamflow declines across the middle East and central Asia with insights from change point detection.
    Sci Rep (2025). https://doi.org/10.1038/s41598-025-32722-3Download citationReceived: 29 August 2025Accepted: 11 December 2025Published: 19 December 2025DOI: https://doi.org/10.1038/s41598-025-32722-3Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative
    KeywordsChange pointClimate changeHuman activitiesPettitt testStreamflow More

  • in

    Satellite altimetry reveals intensifying global river water level variability

    AbstractRiver water levels (RWLs) are fundamental to hydrology, water resource management, and disaster mitigation, yet the majority of the world’s rivers remain ungauged. Here, using 46,993 virtual stations from Sentinel-3A/B altimetry (2016‒2024), we present a global assessment of RWL variability. We find a median global fluctuation of 3.76 m, with pronounced spatial patterns: significant RWL declines across Central North/South America and Western Siberia, and increases across Africa, Oceania, Eastern/Southern Asia, and Northwestern/Central Europe. Seasonality is intensifying in 68% of basins, as high RWLs become more temporally concentrated. Maximum RWLs are declining by 0.88 cm/yr, while minimum RWLs are rising by 1.43 cm/yr. This convergence is reducing seasonal amplitude globally, with the most pronounced changes in the Americas and Central Africa. These shifts coincide with a recent surge in extreme RWL events, particularly after 2021, signaling growing hydrological instability amid concurrent droughts and floods. Our findings underscore the urgent need for adaptive water management in response to accelerating climate pressures.

    Data availability

    Sentinel-3 altimetry data can be downloaded from the Copernicus Data Space Ecosystem at https://dataspace.copernicus.eu/explore-data. Surface Water and Ocean Topography Mission River Database (SWORD) version 15 is available at https://www.swordexplorer.com/. Access to the ancillary data is linked as follows: (1) HydroBASINS: https://data.apps.fao.org/catalog/dataset/7707086d-af3c-41cc-8aa5-323d8609b2d1. (2) In-situ measurements: Water levels and discharge from USGS are available at https://waterdata.usgs.gov/nwis/current/?type=dailystage&group_key=NONE&site_no_name_select=station_nm; Discharge observations from GRDC are accessible via https://portal.grdc.bafg.de/applications/public.html?publicuser=PublicUser#dataDownload/Stations. (3) HydroRIVERS: https://www.hydrosheds.org/products/hydrorivers. (4) GeoDAR: https://zenodo.org/records/6163413. (5) Global Precipitation Climatology Project (GPCP): https://climatedataguide.ucar.edu/climate-data/gpcp-monthly-global-precipitation-climatology-project. (6) GISS Surface Temperature Analysis (GISTEMP): https://data.giss.nasa.gov/gistemp/. (7) Connectivity Status Index: https://doi.org/10.6084/m9.figshare.7688801.v1. Process and result files encoded in JSON format79, including the river water level dataset and the validation dataset, are uploaded at https://doi.org/10.5281/zenodo.17130396. A technical documentation detailing the contents of these result files is also available through this link. Source data, recording all stations’ results, are also provided in this paper. Source data are provided with this paper.
    Code availability

    Water level retrieval from Sentinel-3 altimetry data was performed using MATLAB R2021b, while data postprocessing, including figure plotting, was conducted using Jupyter Notebook (Python 3.10). The code used for processing the results and developing the figures is available at https://github.com/Fangchq/Satellite-rivers/tree/master. The source scripts of our algorithm, available at https://github.com/Fangchq/An-improved-waveform-retracking-method/tree/master, have been slightly tuned to enhance their applicability to global rivers.
    ReferencesLuo, S. et al. Satellite laser altimetry reveals a net water mass gain in global lakes with spatial heterogeneity in the early 21st Century. Geophys. Res. Lett. 49, e2021GL096676 (2022).
    Google Scholar 
    Ntona, M. M., Busico, G., Mastrocicco, M. & Kazakis, N. Modeling groundwater and surface water interaction: An overview of current status and future challenges. Sci. Total Environ. 846, 157355 (2022).
    Google Scholar 
    Scanlon, B. R. et al. Global water resources and the role of groundwater in a resilient water future. Nat. Rev. Earth Environ. 4, 87–101 (2023).
    Google Scholar 
    Quang, N. H., Viet, T. Q., Thang, H. N. & Hieu, N. T. D. Long-term water level dynamics in the Red River basin in response to anthropogenic activities and climate change. Sci. Total Environ. 912, 168985 (2024).
    Google Scholar 
    Wongchuig, S. et al. Multi-Satellite Data Assimilation for Large-Scale Hydrological-Hydrodynamic Prediction: Proof of Concept in the Amazon Basin. Water Resour. Res. 60, e2024WR037155 (2024).
    Google Scholar 
    Acharya, B. S. et al. Unmanned Aerial Vehicles in Hydrology and Water Management: Applications, Challenges, and Perspectives. Water Resour. Res. 57, e2021WR029925 (2021).
    Google Scholar 
    Fang, C. et al. Improving image-based water-level monitoring by coupling water-line detection techniques and the segment anything model. Environ. Model. Softw. 192, 106566 (2025).
    Google Scholar 
    Huang, Q. et al. An improved approach to monitoring Brahmaputra River water levels using retracked altimetry data. Remote Sens. Environ. 211, 112–128 (2018).
    Google Scholar 
    Jiang, L., Nielsen, K., Dinardo, S., Andersen, O. B. & Bauer-Gottwein, P. Evaluation of Sentinel-3 SRAL SAR altimetry over Chinese rivers. Remote Sens. Environ. 237, 111546 (2020).
    Google Scholar 
    Huang, Q., Long, D., Han, Z. & Han, P. High-resolution satellite images combined with hydrological modeling derive river discharge for headwaters: A step toward discharge estimation in ungauged basins. Remote Sens. Environ. 277, 113030 (2022).
    Google Scholar 
    Cooley, S. W., Ryan, J. C. & Smith, L. C. Human alteration of global surface water storage variability. Nature 591, 78–81 (2021).
    Google Scholar 
    Jiang, L., Nielsen, K., Andersen, O. B. & Bauer-Gottwein, P. A Bigger Picture of how the Tibetan Lakes Have Changed Over the Past Decade Revealed by CryoSat-2 Altimetry. J. Geophys. Res.: Atmos. 125, e2020JD033161 (2020).
    Google Scholar 
    Busker, T. et al. A global lake and reservoir volume analysis using a surface water dataset and satellite altimetry. Hydrol. Earth Syst. Sci. 23, 669–690 (2019).
    Google Scholar 
    Papa, F. et al. Ganga-Brahmaputra river discharge from Jason-2 radar altimetry: An update to the long-term satellite-derived estimates of continental freshwater forcing flux into the Bay of Bengal. J. Geophys. Res.: Oceans 117, C11021 (2012).
    Google Scholar 
    Coss, S. et al. Global River Radar Altimetry Time Series (GRRATS): new river elevation earth science data records for the hydrologic community. Earth Syst. Sci. Data 12, 137–150 (2020).
    Google Scholar 
    Huang, Q. et al. Validation and application of water levels derived from Sentinel-3A for the Brahmaputra River. Sci. China Technol. Sci. 62, 1760–1772 (2019).
    Google Scholar 
    Kleinherenbrink, M., Naeije, M., Slobbe, C., Egido, A. & Smith, W. The performance of CryoSat-2 fully-focussed SAR for inland water-level estimation. Remote Sens. Environ. 237, 111589 (2020).
    Google Scholar 
    Chen, J. et al. Evaluation of Sentinel-3A altimetry over Songhua River Basin. J. Hydrol. 618, 129197 (2023).
    Google Scholar 
    Biancamaria, S. et al. Validation of Jason-3 tracking modes over French rivers. Remote Sens. Environ. 209, 77–89 (2018).
    Google Scholar 
    Dinardo, S. et al. Coastal SAR and PLRM altimetry in the German Bight and the West Baltic Sea. Adv. Space Res. 62, 1371–1404 (2018).
    Google Scholar 
    Zhang, X. et al. On the Performance of Sentinel-3 Altimetry Over New Reservoirs: Approaches to Determine Onboard A Priori Elevation. Geophys. Res. Lett. 47, e2020GL088770 (2020).
    Google Scholar 
    Andreadis, K. M. et al. A First Look at River Discharge Estimation From SWOT Satellite Observations. Geophys. Res. Lett. 52, e2024GL114185 (2025).
    Google Scholar 
    Normandin, C. et al. First results of the surface water ocean topography (SWOT) observations to rivers elevation profiles in the Cuvette Centrale of the Congo Basin. Front. Remote Sens. 5, 1466695 (2024).
    Google Scholar 
    Aublanc, J. et al. Sentinel-3 Altimetry Thematic Products for Hydrology, Sea Ice and Land Ice. Sci. Data 12, 714 (2025).
    Google Scholar 
    Zhao, Y., Jiang, L., Zhang, X. & Liu, J. Tracking River’s Pulse From Space: A Global Analysis of River Stage Fluctuations. Geophys. Res. Lett. 50, e2023GL106399 (2023).
    Google Scholar 
    Fang, C. et al. Improved Water Level Retrieval in Complex Riverine Environments: Sentinel-3 and Sentinel-6 Altimetry Over China’s Rivers. Water Resour. Res. 61, e2024WR039705 (2025).
    Google Scholar 
    Wang, H. et al. Anthropogenic climate change has influenced global river flow seasonality. Science 383, 1009–1014 (2024).
    Google Scholar 
    Tollefson, J. How hot will Earth get by 2100?. Nature 580, 443–445 (2020).
    Google Scholar 
    IPCC. Climate Change 2021 – The Physical Science Basis: Working Group I Contribution to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. https://doi.org/10.1017/9781009157896 (2023).Robinson, A., Lehmann, J., Barriopedro, D., Rahmstorf, S. & Coumou, D. Increasing heat and rainfall extremes now far outside the historical climate. npj Clim. Atmos. Sci. 4, 45 (2021).
    Google Scholar 
    Bourgoin, C. et al. Human degradation of tropical moist forests is greater than previously estimated. Nature 631, 570–576 (2024).
    Google Scholar 
    Espinoza, J.-C., Marengo, J. A., Schongart, J. & Jimenez, J. C. The new historical flood of 2021 in the Amazon River compared to major floods of the 21st century: Atmospheric features in the context of the intensification of floods. Weather Clim. Extremes 35, 100406 (2022).
    Google Scholar 
    Huang, A. T., Gillett, Z. E. & Taschetto, A. S. Australian Rainfall Increases During Multi-Year La Niña. Geophys. Res. Lett. 51, e2023GL106939 (2024).
    Google Scholar 
    Aðalgeirsdóttir, G. et al. Glacier Changes in Iceland From ∼1890 to 2019. Front. Earth Sci. 8, 523646 (2020).
    Google Scholar 
    World Weather Attribution. Climate change, not El Niño, main driver of exceptional drought in highly vulnerable Amazon River Basin. (2024).Miller, P. W. & Hiatt, M. Hydrometeorological Drivers of the 2023 Louisiana Water Crisis. Geophys. Res. Lett. 51, e2024GL108545 (2024).
    Google Scholar 
    Neto, J. B. F. et al. Impact of April and May, 2024 extreme precipitation on flooding in Rio Grande do Sul, Brazil. Urban Clim. 61, 102487 (2025).
    Google Scholar 
    Liu, Y., Cai, W., Zhang, Y., Lin, X. & Li, Z. Near-term projection of Amazon rainfall dominated by phase transition of the Interdecadal Pacific Oscillation. npj Clim. Atmos. Sci. 7, 46 (2024).
    Google Scholar 
    Haghtalab, N., Moore, N., Heerspink, B. P. & Hyndman, D. W. Evaluating spatial patterns in precipitation trends across the Amazon basin driven by land cover and global scale forcings. Theor. Appl. Climatol. 140, 411–427 (2020).
    Google Scholar 
    Government of Brazil. Southern Brazil has seen an increase of up to 30% in average annual rainfall over the last three decades. (2024).Perkins-Kirkpatrick, S. et al. Extreme terrestrial heat in 2023. Nat. Rev. Earth Environ. 5, 244–246 (2024).
    Google Scholar 
    Garrido-Perez, J. M. et al. Examining the outstanding Euro-Mediterranean drought of 2021–2022 and its historical context. J. Hydrol. 630, 130653 (2024).
    Google Scholar 
    Copernicus Climate Change Service. Precipitation, relative humidity and soil moisture for May 2024 (2024).World Meteorological Organization. State of Global Water Resources Report 2023. (2024).Espinoza, J.-C. et al. The new record of drought and warmth in the Amazon in 2023 related to regional and global climatic features. Sci. Rep. 14, 8107 (2024).
    Google Scholar 
    CNN. Stark before-and-after pictures reveal dramatic shrinking of major Amazon rivers. (2024).Palmer, M. & Ruhi, A. Linkages between flow regime, biota, and ecosystem processes: Implications for river restoration. Science 365, eaaw2087 (2019).
    Google Scholar 
    Dethier, E. N., Sartain, S. L., Renshaw, C. E. & Magilligan, F. J. Spatially coherent regional changes in seasonal extreme streamflow events in the United States and Canada since 1950. Sci. Adv. 6, eaba5939 (2020).
    Google Scholar 
    Li, L., Long, D., Wang, Y. & Woolway, R. I. Global dominance of seasonality in shaping lake-surface-extent dynamics. Nature 642, 361–368 (2025).
    Google Scholar 
    Papa, F. et al. Water Resources in Africa under global change: monitoring surface waters from space. Surv. Geophys. 44, 43–93 (2023).
    Google Scholar 
    Van Dijk, A. et al. Global water monitor 2024, summary report. (2025).Feng, D. & Gleason, C. J. More flow upstream and less flow downstream: The changing form and function of global rivers. Science 386, 1305–1311 (2024).
    Google Scholar 
    Schmutz, S. & Sendzimir, J. Riverine ecosystem management: science for governing towards a sustainable future. (Springer Nature, 2018).Grill, G. et al. Mapping the world’s free-flowing rivers. Nature 569, 215–221 (2019).
    Google Scholar 
    Han, J. et al. Streamflow seasonality in a snow-dwindling world. Nature 629, 1075–1081 (2024).
    Google Scholar 
    Schwatke, C., Dettmering, D., Bosch, W. & Seitz, F. DAHITI – an innovative approach for estimating water level time series over inland waters using multi-mission satellite altimetry. Hydrol. Earth Syst. Sci. 19, 4345–4364 (2015).
    Google Scholar 
    Santos da Silva, J. et al. Water levels in the Amazon basin derived from the ERS 2 and ENVISAT radar altimetry missions. Remote Sens. Environ. 114, 2160–2181 (2010).
    Google Scholar 
    Huang, Q., Long, D., Du, M., Han, Z. & Han, P. Daily Continuous River Discharge Estimation for Ungauged Basins Using a Hydrologic Model Calibrated by Satellite Altimetry: Implications for the SWOT Mission. Water Resour. Res. 56, e2020WR027309 (2020).
    Google Scholar 
    Stammer, D. & Cazenave, A. Satellite altimetry over oceans and land surfaces. (CRC press, 2017).Altenau, E. H. et al. The Surface Water and Ocean Topography (SWOT) Mission River Database (SWORD): A Global River Network for Satellite Data Products. Water Resour. Res. 57, e2021WR030054 (2021).
    Google Scholar 
    Allen, G. H. & Pavelsky, T. M. Global extent of rivers and streams. Science 361, 585–588 (2018).
    Google Scholar 
    Yamazaki, D. et al. MERIT Hydro: A High-Resolution Global Hydrography Map Based on Latest Topography Dataset. Water Resour. Res. 55, 5053–5073 (2019).
    Google Scholar 
    Lehner, B. & Grill, G. Global river hydrography and network routing: baseline data and new approaches to study the world’s large river systems. Hydrol. Process. 27, 2171–2186 (2013).
    Google Scholar 
    Wang, J. et al. GeoDAR: georeferenced global dams and reservoirs dataset for bridging attributes and geolocations. Earth Syst. Sci. Data 14, 1869–1899 (2022).
    Google Scholar 
    Schneider, D. P., Deser, C., Fasullo, J. & Trenberth, K. E. Climate Data Guide Spurs Discovery and Understanding. Eos, Trans. Am. Geophys. Union 94, 121–122 (2013).
    Google Scholar 
    Lenssen, N. et al. A NASA GISTEMPv4 Observational Uncertainty Ensemble. J. Geophys. Res.: Atmos. 129, e2023JD040179 (2024).
    Google Scholar 
    Birkett, C. M. Contribution of the TOPEX NASA Radar Altimeter to the global monitoring of large rivers and wetlands. Water Resour. Res. 34, 1223–1239 (1998).
    Google Scholar 
    Calmant, S., Crétaux, J.-F. & Rémy, F. Principles of Radar Satellite Altimetry for Application on Inland Waters. (Elsevier, 2016).Biancamaria, S. et al. Satellite radar altimetry water elevations performance over a 200m wide river: Evaluation over the Garonne River. Adv. Space Res. 59, 128–146 (2017).
    Google Scholar 
    Yu, D. et al. Robust estimation and confidence interval in meta-regression models. Comput. Stat. Data Anal. 129, 93–118 (2019).
    Google Scholar 
    Beaton, A. E. & Tukey, J. W. The Fitting of Power Series, Meaning Polynomials, Illustrated on Band-Spectroscopic Data. Technometrics 16, 147–185 (1974).
    Google Scholar 
    Halicki, M., Schwatke, C. & Niedzielski, T. The impact of the satellite ground track shift on the accuracy of altimetric measurements on rivers: A case study of the Sentinel-3 altimetry on the Odra/Oder River. J. Hydrol. 617, 128761 (2023).
    Google Scholar 
    Li, X. et al. Ice thickness and water level estimation for ice-covered lakes with satellite altimetry waveforms and backscattering coefficients. Cryosphere 17, 349–369 (2023).
    Google Scholar 
    Le Gac, S., Boy, F., Blumstein, D., Lasson, L. & Picot, N. Benefits of the Open-Loop Tracking Command (OLTC): Extending conventional nadir altimetry to inland waters monitoring. Adv. Space Res. 68, 843–852 (2021).
    Google Scholar 
    Taburet, N. et al. S3MPC: Improvement on Inland Water Tracking and Water Level Monitoring from the OLTC Onboard Sentinel-3 Altimeters. Remote Sens. 12, 3055 (2020).
    Google Scholar 
    Cerbelaud, A. et al. Peak Flow Event Durations in the Mississippi River Basin and Implications for Temporal Sampling of Rivers. Geophys. Res. Lett. 51, e2024GL109220 (2024).
    Google Scholar 
    Fang, C., Yuan, G., Zheng, Z., Zhong, Q. & Duan, K. Technical note: Monitoring discharge of mountain streams by retrieving image features with deep learning. Hydrol. Earth Syst. Sci. 28, 4085–4098 (2024).
    Google Scholar 
    Sharannya, T. M. & Mahesha, A. In Modern River Science for Watershed Management: GIS and Hydrogeological Application (eds S. Satheeshkumar, V. Thirukumaran, & D. Karunanidhi) 59–69 (Springer Nature Switzerland, 2024).Fang, C. & Long, D. Satellite altimetry reveals intensifying global river water level variability. Zenodo. https://doi.org/10.5281/zenodo.17130396 (2025)Download referencesAcknowledgementsThis study was supported by the National Natural Science Foundation of China (Grant No. 52325901 to D.L. and Grant no. 42571432 to Q.H.). F.P. is supported by the CNES-TOSCA SWOT Science Team Project SAMBA. This work was partially funded by CNES (Centre National d’Etudes Spatiales) under the framework of the TOSCA program. The European Space Agency (ESA) is acknowledged for providing Sentinel-3 altimetry data, while in-situ measurements were made available by the United States Geological Survey (USGS), the Global Runoff Data Center (GRDC), and the Information Center of the Ministry of Water Resources of the People’s Republic of China.Author informationAuthor notesThese authors contributed equally: Chenqi Fang, Di Long.Authors and AffiliationsDepartment of Hydraulic Engineering, State Key Laboratory of Hydroscience and Engineering, Tsinghua University, Beijing, ChinaChenqi Fang, Di Long & Huaichuan LiuKey Laboratory of Hydrosphere Sciences of the Ministry of Water Resources, Tsinghua University, Beijing, ChinaChenqi Fang, Di Long & Huaichuan LiuSchool of Water Resources and Environment, China University of Geosciences, Beijing, ChinaQi HuangUniversité de Toulouse, LEGOS (IRD/CNRS/CNES/UT), Toulouse, FranceJean-François Cretaux & Fabrice PapaInstitut de Recherche pour le Développement, Universidade de Brasília, Brasília, BrazilFabrice PapaISPA, UMR1391 INRAE/Bordeaux Sciences Agro, Villenave d’Ornon, FranceFrédéric FrappartDepartment of Civil and Environmental Engineering, University of Massachusetts Amherst, Amherst, USAColin J. GleasonChina Institute of Water Resources and Hydropower Research, Beijing, ChinaChunhong HuAuthorsChenqi FangView author publicationsSearch author on:PubMed Google ScholarDi LongView author publicationsSearch author on:PubMed Google ScholarQi HuangView author publicationsSearch author on:PubMed Google ScholarJean-François CretauxView author publicationsSearch author on:PubMed Google ScholarFabrice PapaView author publicationsSearch author on:PubMed Google ScholarFrédéric FrappartView author publicationsSearch author on:PubMed Google ScholarHuaichuan LiuView author publicationsSearch author on:PubMed Google ScholarColin J. GleasonView author publicationsSearch author on:PubMed Google ScholarChunhong HuView author publicationsSearch author on:PubMed Google ScholarContributionsD.L., C.H., and C.F. developed the concept and methodology of this study. C.F., D.L., Q.H., and H.L. performed the data processing and analysis with support from all other authors. D.L., C.F., Q.H., J.F.C., F.P., F.F., H.L., C.J.G., and C.H. discussed the results and improved the writing of this manuscript.Corresponding authorCorrespondence to
    Di Long.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Peer review

    Peer review information
    Nature Communications thanks the anonymous reviewers for their contribution to the peer review of this work. A peer review file is available.

    Additional informationPublisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary informationSupplementary InformationTransparent Peer Review fileSource dataSource DataRights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleFang, C., Long, D., Huang, Q. et al. Satellite altimetry reveals intensifying global river water level variability.
    Nat Commun (2025). https://doi.org/10.1038/s41467-025-67682-9Download citationReceived: 09 June 2025Accepted: 04 December 2025Published: 19 December 2025DOI: https://doi.org/10.1038/s41467-025-67682-9Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Tariff familiarity sustains household water conservation

    AbstractDue to rapid urbanization and income growth, residential water consumption worldwide is increasing much faster than in other sectors. This exponential growth in demand threatens freshwater resources, and policymakers across the globe are attempting to curb usage by redesigning tariff structures and implementing price hikes. Using a unique seven-year dataset of daily household water usage from multiple counties in China, we show that tariff reforms that simultaneously change both structure and price yield only short-lived conservation effects, as usage rebounds within months. Instead, our empirical results confirm that staggered tariff reforms—introducing price hikes only after households have adapted to a new structure—can reduce water consumption. Policy simulations suggest that a staggered reform approach could achieve an additional 4-percentage-point reduction in water usage. These findings provide new evidence on how tariff design shapes consumption behaviour and offer actionable, evidence-based insights for policymakers seeking to design effective, equitable, and sustainable water management strategies.

    Similar content being viewed by others

    Tracking and tracing water consumption for informed water sensitive intervention through machine learning approach

    Article
    Open access
    09 April 2024

    The effects of policy announcement, prices and subsidies on water consumption

    Article

    30 January 2023

    Long-term water conservation is fostered by smart meter-based feedback and digital user engagement

    Article
    Open access
    07 May 2021

    IntroductionIt has been highlighted recently in a global report that the hydrological cycle is out of balance for the first time in human history, and one of the main causes is overconsumption1. The largest water users globally are high-income countries, and low- and middle-income countries (LMICs) with large population (Fig. 1). More importantly, due to urbanization and economic expansion, municipal water consumption has grown at more than twice the rate of any other sector over the past decade2 (Fig. 1), although agriculture continues to account for the majority of global freshwater withdrawals (Fig. S1). In contexts marked by rapid urban growth, limited supply, and weak governance, unmanaged household water demand can considerably contribute to local water scarcity.Fig. 1: Global annual freshwater withdrawals.a The annual freshwater withdrawals by country in 2022. b–g The percentage increase relative to reference year in freshwater usage by household (blue solid), agriculture (gray solid) and industry (gray short dash) sector. Data from (Food and Agriculture Organization of the United Nations – AQUASTAT, https://data.apps.fao.org/aquastat/?lang=en).Full size imageWhile increased residential water use may reflect expanded access in some regions, such aggregate trends often obscure persistent inequalities in water distribution. Recent studies show that water consumption is frequently concentrated among high-income households3, even as large segments of the low-income urban residents remain unconnected to safe piped water4. This trend is likely to continue, and even accelerate in LMICs, posing a serious challenge to achieving Sustainable Development Goal (SDG) 6 – ensuring universal and equitable access to clean water2,5. Effectively curbing household water usage is therefore essential to closing the demand-supply gap and realizing this global target.China exemplifies this challenge, where rapid economic growth has been closely linked to increasing water scarcity. Currently, over 700 million Chinese experience water shortages for at least one month each year6. This crisis is projected to intensify for two key reasons. First, for every 1% increase in China’s urbanization rate, household water consumption rises by 1.67 billion cubic meters7. Second, climate change is projected to drive household water demand up by 3.1–10.4% by 2030–2049 and by 7.2–43.8% by 2080–20998. Taken together, residential water consumption in China is projected to increase by 15.4 billion cubic meters (a 25.6% rise from 2023 levels) by 2030 under the Shared Socioeconomic Pathway 245 scenario. By then, household water demand is projected to surpass industrial usage, further straining water resources and exacerbating the country’s water scarcity crisis7.To address this impending crisis, the Chinese government has mostly focused on supply-side policies (e.g., desalination, regional water transfers) to meet the ever-growing water demand. However, these measures alone are insufficient as a comprehensive solution requires demand-side management as well, particularly through water pricing reform and urban water conservation initiatives5. Toward this end, a new report issued by the OECD emphasized using water tariff reforms to better manage usage9.From a sustainable development perspective, water pricing reform is a powerful tool for balancing efficiency and equity5. By adjusting both the price levels and tariff structures, such as rising marginal prices and subsidized lifeline blocks, policymakers can incentivize conservation while ensuring equitable access to water. Moreover, artificially low water tariffs not only encourage discretionary use but also constrain the financial capacity of utilities, diverting resources away from investments needed to expand access equitably4. In this light, tariff reform plays a critical role in achieving SDG 6—not only by curbing inefficient or excessive residential water consumption, but also by enhancing the financial sustainability of water utilities and enabling service expansion to marginalized households.However, implementing water pricing reform is challenging. In practice, while simultaneous adjustment of water prices and tariff design appear to yield the best of both worlds (e.g., efficiency and equity), consumers frequently respond in irrational or unanticipated ways when confronted with complex pricing schemes. One explanation is that limited attention arises because water expenses typically constitute only a small share of total household expenditures, causing consumers to invest insufficient effort in fully understanding their bills10. Information asymmetry also contributes to misunderstandings of tariff structures. Many residential water tariffs employ increasing block tariffs (IBTs), characterized by tiered thresholds and progressively higher marginal prices. Consumers typically are unaware of the marginal price they face and frequently overestimate it11, which undermines their ability to make informed consumption decisions.These cognitive biases lead households to rely on heuristic decision-making rather than fully optimized choices when making water usage decision. A recent study showed that residential water customers with heuristic decision rules make water consumption decision based on easily-accessible bill information rather than detailed tariff structures12.If households anchor their perceptions on inflated prior bills, they may initially overreact to a transition from flat rates to IBTs, perceiving water as more expensive than it actually is. Conversely, upon discovering that water is cheaper-than-expected, consumers may paradoxically increase their usage over time. Thus, an “effective” water tariff reform should meet three key criteria: it should be easily understood by the general public, clearly communicates water scarcity, and encourages sustained behavioral changes aligned with long-term conservation goals (e.g., exceeding two years13).In this regard, a key knowledge gap in “pricing water right” is whether we should reform the tariff design and price levels concurrently or separately. Even though there is a voluminous literature examining how households react toward water pricing14,15,16,17, few if any, have attempted a simultaneous examination of various types of tariff reforms. Moreover, evidence is especially scarce in LMICs, where institutional capacity, billing transparency, and public comprehension vary widely. China, with its diverse reform history, offers a natural laboratory for testing these questions.Water pricing reform in China began in the late 1970s under a “cost-reflected tariff” scheme that considered only supply costs (e.g., raw water extraction) while excluding investment profits, wastewater fees, and resource charges. Tariffs during this period averaged less than 0.05 CNY m−3, resulting in extremely low prices and poor efficiency (see Table S1 for a list of policy documents on China’s water pricing policies, and Fig. 2 for overview of policy reforms timeline). Reforms in the 1990s and early 2000s progressively incorporated wastewater treatment fees and water resource charges, laying the foundation for two major innovations: increases in base water prices to reflect production and environmental costs, and the nationwide promotion of IBTs to replace flat-rate structures. Under the 2002 directive, pilot cities were required to adopt three-tier tariff systems covering 80% of average use in the first block and maintaining at least a 1:1.5:3 price ratio across tiers. Localities facing severe scarcity were encouraged to adopt steeper ratios to enhance conservation. Implementation, however, proceeded unevenly. While early adopters such as Shenzhen transitioned in the 1990s, many counties delayed adoption until after 2010 because of technical constraints (e.g., inadequate metering) and inter-agency negotiations. Subsequent regulations in 2015 further clarified wastewater fee structures, mandating minimum rates across urban and rural jurisdictions. As a result, by the 2010s most counties had implemented multi-tier tariffs, but substantial heterogeneity remained in timing, structure, and communication strategies. This diversity in reform experiences—spanning both shifts from flat to tiered tariffs and price adjustments within existing tiers—creates a unique quasi-experimental setting for identifying the causal effects of tariff reform on household demand.Fig. 2: Timeline of China’s water tariff policy reforms.Grey shade indicates flat water tariff across China, light green indicates pilot implementation of IBT, and dark green indicates full adoption of IBT.Full size imageIn this work, we combine a seven-year panel of household-level daily water-usage data from 25 counties across seven provinces (2012–2019) with detailed administrative records on tariff reforms to estimate how different reform types influence residential consumption (see section 4.2 on Dataset). We make three key contributions to the literature. First, we provide one of the earliest causal estimates of household water-demand responses using high-frequency administrative data, exploiting quasi-experimental variation in the timing and nature of tariff changes across counties17,18,19,20,21,22. Second, we address a previously unexplored policy dilemma—whether tariff design and price levels should be reformed concurrently or sequentially—by systematically contrasting their short- and long-term effects18,23,24. Third, we extend the empirical evidence base to one of the most water-stressed and policy-active settings in the Global South25,26, offering generalizable insights for LMICs seeking to balance efficiency, equity, and fiscal sustainability in water governance1,27,28,29,30.We show that when tariff reforms simultaneously modify both structure and rates, households initially reduce usage by about 6% but gradually rebound, maintaining only a 1.5% long-term reduction. In contrast, when only rates change under familiar tariff structures, usage declines by 5–6.5% and remains persistently lower even after two years. Through a series of robustness tests—including variations in communication strategy, tariff complexity, and survey-based assessments of billing comprehension—we demonstrate that tariff familiarity enhances the durability of conservation by allowing consumers to correctly internalize price signals. Simulations indicate that applying this insight could reduce household water use by 5.4%, compared with only 1.4% under current mixed reform practices. Our findings reveal that sequencing and clarity in tariff design are as crucial as price magnitude for sustaining conservation, offering actionable guidance for designing effective and equitable water-pricing reforms.ResultsBaseline resultsWe begin analysis by examining the overall impacts of water tariffs on household water usage for the two types of tariff reforms. The results in Table 1 show that within-tier (“within-IBT”) changes have the largest impact on household water usage at an average 6% (P < 0.001, 95% CI [−0.07, −0.03]) decrease. The change from flat to increasing block tariffs (“flat-to-IBT”) also causes a statistically significant decrease but its effects are much smaller at 1.5% (P = 0.099, 95% CI [−0.03, 0.00]).Table 1 Baseline resultsFull size tableWhile the initial set of results provides us an overall view of the relative effectiveness of each type of tariff change on water usage, we can further exploit the granularity of our dataset to observe the evolution of households’ responses over time. To do that, we restricted the initial sample to one month after tariff reforms, and gradually expanded the dataset to include more months.We begin with “flat-to-IBT” reforms in Fig. 3a. Compared to households located in counties with flat tariffs, water usage decreased by about 6% (P = 0.012, 95% CI [−0.10, −0.01]) immediately (i.e., one month) after the policy change. However, this behavioral change is not fully sustained as there is a gradual rebound in household usage over time. Eventually, households’ water usage decreased by only 1.5% in the long-term as shown earlier. This dichotomy in impacts over time also reflects the mixed evidence observed from earlier studies31,32. For instance, one study found that “flat-to-IBT” reform reduced annual residential usage by 3–4% in the short-run and 5% in the longer run31. However, another study found that the same reform did not reduce usage31. These two contrasting findings imply that households may respond differently to the “flat-to-IBT” reform in the short and long term.Fig. 3: Baseline impact of tariff reforms.a The baseline impact of tariff reforms on log-transformed daily household water usage of households. In each regression, we begin by using data up until one month after reform, and gradually add observations from latter time periods. b, c The parallel trend tests, which is implemented by including an interaction term for each time period (i.e., days relative to reform) with the treatment or reform variable. In each graph, the dots indicate coefficient size, and the bars represent 95% confidence intervals. The regression models are estimated with daily weather controls (temperature, total precipitation, relative humidity, wind speed, hours of sunlight, and atmospheric pressure), calendar date fixed-effects, and households fixed-effects. Standard errors are clustered two-way at the county, and year-by-month level.Full size imageThe second reform we examine is “within-IBT” change. In this scenario, households were already under an IBT structure, and the price change they experienced is a direct change to volumetric rates within each usage block. Figure 3a shows water usage decreased by around 5% (P = 0.002, 95% CI [−0.07, −0.02]) one month after rates changed. Unlike the “flat-to-IBT” scenario, the reduced usage is largely sustained over time. Specifically, household water consumption decreased by around 7.0% (P = 0.004, 95% CI [−0.112, 0.028]) two months later, and stabilized at approximately 6% in the longer term.A central assumption of the two-way fixed-effects model is that treatment and comparison groups share common trends before policy implementation, known as the parallel trends assumption. We test this by interacting treatment status with time dummies for each day relative to the benchmark, 15 days prior to the reform dates (See SI Section 1 for details on the methods). Figure 3b, c show no significant pre-treatment trends for both of the reforms.One plausible explanation for the results obtained so far is that the magnitudes of price changes may vary substantially, and more importantly, be correlated with the different types of tariff changes. For instance, if “flat-to-IBT” reform price changes are on average smaller than “within-IBT” changes, then it is possible that the relatively smaller long-term effects in the former are due to price change magnitude rather than differences in how consumers perceive tariff structures.As changes to water tariffs are relatively infrequent across most countries and local jurisdictions, most studies can only examine the impact of one-time changes. By contrast, this study enables us to map out different magnitudes of changes as our dataset includes multiple counties that experienced water tariff reforms at different points in time, each with their own unique price levels.To rule out price magnitude effects, we need to first compute marginal water prices of each tariff type. To do so, we categorize each household into their respective tariff rate blocks. For the “flat-to-IBT” reform, we use pre-reform average monthly water consumption of each household (denoted by ({C}_{0})) in the treatment group to assign them their post-reform tiered tariff blocks. For instance, if a household averaged 25 m3 monthly water consumption before “flat-to-IBT” reform and the county where the household is located will have three tariff blocks of 0 to 20 m3, 20 to 30 m3, and more than 30m3, we assign this household to the middle block.Based on ({C}_{0}), we can then derive pre-reform marginal price ({{{{rm{MP}}}}}_{{{{rm{i}}}},0}). For the “flat-to-IBT” sample, the marginal price is equal to the corresponding unit price; for the “within-tier” reform, the marginal price is the unit price for an additional unit of quantity. We then calculate the marginal price after the reform (denoted by ({{{{rm{MP}}}}}_{{{{rm{i}}}},1})), using ({C}_{0}), and the new tariff scheme. Finally, we compute the proportional change in marginal water price (left({{{{rm{Price}}}}_{{{rm{change}}}}}_{{{{rm{i}}}}}right)) for each household.A direct way to assess price magnitude effects is by replacing the binary treatment variable with a log-transformed marginal price change variable. In this manner, we can interpret the treatment coefficient as the elasticity of tariff on water usage (i.e., the percentage change in water demand for a one percentage change in price).However, as prices are potentially endogenous to the household, we use marginal price change computed at each household’s pre-reform average consumption33, which is predetermined and thus plausibly exogenous to post-reform behavior. Another approach is by using changes in the marginal price schedule at each block as the instrumental variable21,34. This two-stage least-squares (2SLS) approach leverages on the fact that the full structure of the block-rate schedule is correlated with the price changes households face but is exogenous to their contemporaneous demand shocks.Figure S2 reports both OLS (Panel a) and 2SLS (Panel b) estimates. Standard diagnostic tests—including the Cragg-Donald F-statistic and Kleibergen-Paap LM statistic—confirm strength of the instrumental variable. The close alignment between OLS and 2SLS estimates further reinforces the robustness and internal validity of our findings. Similar trends persist even after controlling for price magnitude. Households experiencing “flat-to-IBT” change continue to have a reduce-and-rebound effect. In contrast, the “within-IBT” change delivers relatively consistent reduction in water usage over time.An indirect way to assess price magnitude effects is to combine counties that experienced similar magnitudes of price changes and conduct sub-group analyses. According to Fig. S3, we observe results consistent with the baseline’s, regardless of whether the percentage price change was below 10%, or between 10% and 15%.Heterogeneity analysesThe results so far are in part consistent with evidence that households reduce water usage in response to price hikes. However, we also show that, not only do different modalities of tariff reforms have varied impacts on water usage but households take different amount of time to process price effects. In this sub-section, we conduct heterogeneity analyses by prior water usage to examine if households’ sensitivity to price changes is in accordance to the amount of water they use prior to policy changes.To do so, we follow the same procedure outlined in Section 2.1 to assign their water usage blocks.Panels a and b of Fig. S4 show that for the “flat-to-IBT” reform, households in Block 1 (the low-usage group) reduced water consumption by 5% (P = 0.033, 95% CI [−0.10, −0.01]) at the beginning. By contrast, their high-usage counterparts (i.e., Block 2) decreased usage by 14% (P < 0.001, 95% CI [−0.180, −0.101]) in the same period. This disparity is not surprising as “flat-to-IBT” reform is designed to disproportionately restrain high-usage groups. However, a trend common across both groups is that water usage rebounded in later periods and, in this case, the low-usage group fully rebounded in its water consumption by some 18 months after the policy change, while the high-usage group rebounded to an average 8% (P < 0.001, 95% CI [−0.109, −0.050]) decrease over time.We next examine “within-IBT.” Panels a and b of Fig. S4 show similar patterns from earlier where the low-usage group (Block 1) reduce usage by a smaller amount (3%, P = 0.020, 95% CI [−0.06, −0.01]) compared to the high-usage (Block 2) group’s 9% (P = 0.006, 95% CI [−0.15, −0.03]). However, a key difference here is that the decrease in water usage is consistent over time for both groups.The result that high-usage households exhibit greater responsiveness to price changes also suggests that well-designed IBTs may have progressive features by encouraging conservation among higher-consuming households, who are often relatively wealthier3. A more steeply tiered IBT structure—with a lower-priced first-tier and a higher-priced second-tier—could enhance both equity and efficiency by promoting cross-subsidization and targeting discretionary water use.As the vast majority of households (around 90–95%) belonged to the low-usage group, we further examine whether the previous results are replicable within Block 1. We further split low-usage (i.e., Block 1) households into two equal-sized groups based on usage. The results in Fig. S5 are consistent from earlier. For the “flat-to-IBT” change, the higher-users within Block 1 are more sensitive toward tariff changes. However, policy impacts continue to decay over time. For the “within-IBT” changes, decrease in water usage is steady over time.In all, this set of results is consistent with our baseline findings as they show that households can better process “within-IBT” tariff changes.To validate the baseline results, we perform additional analyses (SI Section 2). We first investigate the impact of annual household water budgets17 by excluding counties with such policies from the baseline model. We then conduct a placebo test by randomly assigning treatment groups and treatment timing to ensure the baseline findings are not driven by other factors (Fig. S6). Further robustness checks—including using alternative clustered standard errors, nonlinear effects of weather variables and excluding abnormal water consumption observations—show that the baseline results remain robust (Fig. S7). We also rule out the possibility that other confounding policies that impact water use (such as, water conservation campaigns, water efficiency subsidies, watering restrictions) could coincide with the water tariff (Fig. S8 and Table S2). Lastly, recent studies suggest that the usual two-way fixed-effects (TWFE) model may not accurately estimate policy effects in scenarios of staggered treatment adoption35,36. We reinforce the robustness of our baseline results using estimators proposed by Gardner37 and Callaway and Sant’Anna35, demonstrating that our TWFE model provides reliable estimates of reform effects (Table S3).Possible mechanismsThe results so far suggest that households are less reactive when confronted with a change in tariff structure compared to changes in volumetric rate. Here, we explore several reasons for such behavioral patterns.First, “flat-to-IBT” involves changes to both the tariff structure and the volumetric rate (at least at the higher tiers). We can further isolate these two separate effects as there are some counties that switch from “flat-to-IBT” while keeping rates at the first tier intact. In such cases, since there are no changes to the water bills for households in the first usage tier, they should not exhibit any behavioral changes. Figure 4 separates the “flat-to-IBT” counties into these two categories, and includes only households in first usage tier. While water usage of households in the “structure-only” group rebound much faster than the “structure and price” group, both sets of households have the same reduce-and-rebound usage pattern. From here, we conclude that households inadvertently react to changes in tariff structure even if the underlying tariff did not change for them. We also note that although the corresponding coefficients across these two groups are not statistically different, the larger magnitude observed in the “structure-only” group may be partially explained by their exposure to more substantial second-tier price increases during the “flat-to-IBT” reform. However, this finding merits future research.Fig. 4: Heterogeneity analysis by price change in flat to IBT reform.This figure shows the impact of flat-to-IBT reform on water usage, with counties separated by whether they also experienced price changes. Structure only households are those with average monthly consumption within the first tier of the IBT and located in counties that transitioned from a flat-rate to an IBT structure while keeping first-tier rates unchanged. Structure and price households refer to those in counties where both the tariff structure and price levels were adjusted. In each regression, we begin by using data up until one month after reform, and gradually add observations from latter time periods. The dots indicate coefficient size, and the bars represent 95% confidence intervals. The regression models are estimated with daily weather controls (temperature, total precipitation, relative humidity, wind speed, hours of sunlight, and atmospheric pressure), calendar date fixed-effects, and households fixed-effects. Standard errors are clustered two-way at the county, and year-by-month level.Full size imageSecond, one plausible reason why households undergoing “flat-to-IBT” reforms exhibited reduce-and-rebound effects is that they could not fully comprehend the implications of the tariff change. We test this hypothesis using public announcements of tariff changes, and survey responses. County governments communicated the change in tariff structure to their citizens through public announcements. However, since there is no fixed template, there is substantial variation in how these announcements are phrased. We exploit these heterogeneities by categorizing them into two groups according to their inclusion of useful or pertinent information. Specifically, we classify announcements as effective if they included at least five of the following six elements: concise length (under 800 words), policy start date, old and new tariff rates, tier-specific usage thresholds, and estimated bill changes. For details and coding examples, see SI Section 3, and Table S4. Figure 5 shows that households from “flat-to-IBT” counties that effectively communicated the tariff change reduced water usage in a consistent manner, and did not rebound. On the other hand, households from counties with less effective communication have the same reduce-and-rebound effect we saw earlier. The second method we investigate is via an individual-level survey implemented on 853 respondents located across the country (Fig. 6). We asked them to compute total changes to a water bill under two scenarios: i) “flat-to-IBT” and ii) “within-IBT”. Findings from our survey mirror our empirical findings as only 33.9% of respondents correctly answered how “flat-to-IBT” change will affect water bills. Moreover, the vast majority of responses (59%) overestimated the changes in water bills. On the other hand, 80.8% answered correctly when they were quizzed on the implications of within-IBT changes, and the wrong responses were evenly split between over- and underestimation (9.3% and 9.8%, respectively). As we required respondents to compute hypothetical bill changes, the error rates may be due to questions’ complexity instead of the true misunderstanding of water tariff. In SI section 4 and Table S6, we provide evidence that the mistakes mainly reflect true misunderstanding about the water tariff.Fig. 5: Heterogeneity analysis by flat-to-IBT policy announcements effectiveness.This figure shows the impact of flat-to-IBT water tariff reform on log-transformed daily household water usage, with counties separated by whether their public announcements included useful or pertinent information. We begin by using data up until one month after reform, and gradually add observations from latter time periods. The dots indicate coefficient size, and the bars represent 95% confidence intervals. The regression models are estimated with daily weather controls (temperature, total precipitation, relative humidity, wind speed, hours of sunlight, and atmospheric pressure), calendar date fixed-effects, and households fixed-effects. Standard errors are clustered two-way at the county, and year-by-month level.Full size imageFig. 6: Computation of water bill changes.This figure shows the proportion of survey respondents that answered correctly to how different types of tariff reforms will affect water bills. The bars denote average percentage and the lines refer to the 95% confidence interval of the mean percentage. Number of survey observations: 853. Data is obtained from an online survey sampled across China. Descriptive statistics are collected in Table S5.Full size imageThird, the results so far confirm that households can better comprehend the implications of “within-IBT” reform. We can further test this hypothesis by examining the effects of another straightforward volumetric rate increase based on wastewater treatment fees (WWTF). Water bills in China consist of two components: i) water tariffs and ii) WWTF. While the latter is itemized as a separate bill item for transparency and audit reasons, it is identical to the former in its contribution to the final water bill. In this regard, households should react similarly to an increase in WWTF as they would with “within-IBT” since both are volumetric rates. However, Fig. 7 shows that water usage does not respond to WWTF reforms whether in the short- or long-term. One plausible explanation is that Chinese households mostly do not know that WWTF are part of their water bills, and so did not react to this change. In the same survey (Fig. 8), we find that only 21.1% of respondents were aware their water tariff included a WWTF component even though this tariff structure is ubiquitous nationally. Even when informed that China’s water tariffs were designed with a WWTF component, less than half the respondents (42%) were aware that all residential water usage is subjected to this fee. Instead, most of them (72%) had the wrong impression that only “dirty” water collected from toilets or kitchen were charged under this tariff.Fig. 7: Impacts of change to WWTF on household water usage.This figure plots the impact of change to wastewater treatment fee (WWTF) reform on log-transformed daily household water usage. We begin by using data up until one month after reform, and gradually add observations from latter time periods. The dots indicate coefficient size, and the bars represent 95% confidence intervals. The regression models are estimated with daily weather controls (temperature, total precipitation, relative humidity, wind speed, hours of sunlight, and atmospheric pressure), calendar date fixed-effects, and households fixed-effects. Standard errors are clustered two-way at the county, and year-by-month level.Full size imageFig. 8: Consumers’ knowledge about WWTF.This figure shows the extent to which survey respondents are aware of wastewater treatment fee (WWTF) policies. Objectively correct answers are highlighted in yellow. Number of survey observations: 853. Data is obtained from an online survey sampled across China. Descriptive statistics are collected in Table S5.Full size imageIn all, this series of examinations confirm that households will only behave predictably when water tariff is reformed under familiar circumstances.Policy simulationsOur findings so far confirm that household water usage is most sensitive toward price increase when they are already familiar with the tariff structure or fees components. As such, we conduct two policy simulations using these insights.The first aims to assess the impact on household water usage if Chinese local governments adopted a staggered approach when transitioning from “flat-to-IBT”. In this scenario, the tariff structure would be changed in the first year while retaining the same volumetric rate in the first tier. In the second year, the tiered rates would then be increased to the originally intended levels (see SI Section 5 for full exposition of policy simulation). Compared to the existing scenario where both structure and tariff changed simultaneously, our proposed staggered tariff reform will reduce household water usage by 5.4% compared to 1.4% under current policies – a difference of four percentage points.Similarly, the second simulation aims to assess the impact on household water usage when the increase in wastewater treatment fees is reframed as a volumetric tariff increase. In this scenario, our simulation shows that household water usage reduces by around 5% compared to the current no-effect outcome. This is because households are much more likely to adopt behavioral changes if the tariff increase is expressed in familiar terms.DiscussionSDG 6, which aims to ensure clean water for all, is increasingly under threat due to ever-rising demand, and diminishing supply caused by climate change. Among the various demand drivers, household is the fastest-growing sector due to urbanization and rapid income growth in LMICs2,3,38. China exemplifies this trend, with residential water usage projected to surpass industrial consumption by 2030, making it the second-largest water-consuming sector7. This shift underscores the urgent need for demand management policies. One of the key recommendations by the Global Commission on the Economics of Water is to eliminate water underpricing to better reflect scarcity and incentivize conservation1. While extensive research has explored the effects of water tariffs on consumption, critical knowledge gaps remain on understanding how households react toward tariff reforms. Addressing these gaps is crucial for designing effective policies to balance water demand and ensure long-term sustainability.In particular, a central question emerges for policy makers and utility managers: should reform focus on tariff design, prices, or a combination of both to effectively reduce water usage. Here, we pair a unique dataset of household-day water usage across multiple counties over a seven-year period, spanning the widely-varying policy landscape of China’s water tariff structures to conduct natural experiments on how residents alter their water usage in response to two main types of tariff changes.When the reform involves both structure and price, i.e. flat rate to increasing block tariffs (“flat-to-IBT”), households exhibit a reduce-and-rebound behavioral change as they reduce usage by a large amount (6% decrease) in its immediate aftermath. However, usage rebounds steadily over time to a 1.5% usage decrease in the long-run. This pattern likely reflects learning lags: initially, households may overreact due to anchoring on inflated pre-reform bills and prices. As they gain experience and realize that bills and marginal prices are lower than expected, their consumption adjusts upward.When the change is “within-IBT,” policy impacts are much more consistent as households decrease water usage by around 5–6.5% across short- to long-term.We also exploit other types of tariff changes, heterogeneity in information dissemination by local governments, and survey responses to further explore how households react toward reform in tariff rates vis-à-vis tariff structure. In all, our body of evidence conclude that households are best assessing impact of changes in tariff rates made under a familiar design. In contrast, changes to tariff structure are often confusing for households, and results in muted responses to water usage.This study yields the following policy implications that are generalizable to any locations looking to reform water tariffs.First, decouple tariff structure from price adjustments. There are still 1.8 billion people worldwide without access to safe drinking water39. It is projected that urbanization, and climate change will greatly worsen water scarcity across the world, bringing us further away from SDG 62,3,40. One way to correct the imbalance between supply and demand is to not only “price water right,” but also implement new tariff designs that can ensure higher overall prices without jeopardizing basic affordability. Utility managers often package these two changes together, and our findings show that such policies fail to reduce water usage. On this note, a key insight from our study is for policymakers to decouple reforms to tariff structure from price increase. As much as possible, they should first allow households to familiarize with the new tariff structure before introducing additional rate changes. It is likely that the “structure versus price” dilemma identified here is not unique to the Chinese context, but represents a broader challenge in many LMICs. Nearly 44% of urban water utilities in LMICs still employ flat tariffs41, suggesting many are candidates for flat-to-IBT or other reforms. Moreover, empirical evidence from diverse contexts—including South Africa, Mexico, and Spain—consistently indicates the complexity of IBTs and their poor comprehension by consumers10,42,43.Second and related, if it is only feasible to package tariff structure and rate changes together, the next best solution is to ensure that communications on these reforms are simplified and concise. There is evidence that attempts to improve price salience through increased billing frequency or non-price interventions like social norm comparisons lead to modest and often short-lived, or counterintuitive outcomes—such as increased water usage11,44,45,46,47,48. We find that households residing in counties that communicated effectively on the “flat-to-IBT” changes exhibit consistent behavioral shifts compared to those in counties that failed to do so. These evidence may have clear implications for water rate design. Enhancing price salience—e.g., through simplified billing formats or clearer communication, not just social comparison or increasing water bill frequency—could help ensure that price signals are both understood and sustained over time49. One plausible strategy to ensure effective communications is to conduct multiple focus group discussions to ensure that ordinary residents can readily understand these changes50,51.Third, reframe obscure charges using more intuitive or familiar language. From the perspective of price transparency, it is good that water bills are itemized into distinctive components. However, this also creates a new problem where households may not readily associate with some obscure components, and thus fail to curb water usage. According to mental accounting framework52, households may categorize WWTF as a separate charge distinct from the water tariff, because it is itemized differently and often framed in unfamiliar terminology. To improve the behavioral effectiveness of such rate components, utilities might consider reframing them in ways that emphasize their impact on the overall bill. For instance, instead of announcing an increase in WWTF alone, it may be more effective to present the change as an increase in total water bills.In summary, households’ water usage is projected to rise exponentially across much of the LMICs in the near future40,53. An inevitable policy solution is to increase water tariffs to curb the rising demand. While China has mostly wasted this opportunity in maximizing the effectiveness of using tariff reforms to reduce water usage, lessons from them are valuable to the rest of the LMICs embarking on similar reforms. Policy simulations conducted using our findings show that by simply staggering tariff reforms, and reframing tariff components in terms familiar to consumers can bring about additional reduction in water usage of 4 to 5 percentage points.MethodsEstimating equationThe water tariff reform examined in this study constitutes a quasi-natural experiment with different counties implementing reforms at varying times. Consequently, we employ a two-way fixed effects (TWFE) model to estimate the effect of the water tariff reform on household water usage. This method is particularly well-suited for evaluating staggered policy rollouts and has been widely used in related empirical work on environmental and public policy evaluation54,55,56 (see SI Section 6 for more discussion on policy evaluation methods). All analyses were performed using STATA/MP 16, and the effect of each water tariff reform on household water usage is estimated using the following TWFE model:$${{mathrm{ln}}}left({{{{rm{usage}}}}}_{{{{rm{ict}}}}}right)={beta }_{0}+beta times {{{{rm{reform}}}}}_{{{{rm{ct}}}}}+{{{{bf{W}}}}}_{{{{rm{ct}}}}}+{lambda }_{{{{rm{t}}}}}+{gamma }_{{{{rm{i}}}}}{+varepsilon }_{{{{rm{it}}}}}$$
    (1)
    where ({{mathrm{ln}}}left({{{{rm{usage}}}}}_{{{{rm{ict}}}}}right)) represents log-transformed water usage of the household (i) at date (t) in county (c).({{{{rm{reform}}}}}_{{{{rm{ct}}}}}) indicates whether county c where household i is located has already implemented water tariff reform at date t. Here we primarily consider two types of reforms: i) flat-to-tiered tariffs, ii) changes in volumetric water price within tiered tariffs.({{{{bf{W}}}}}_{{ct}}) is a vector of weather control variables which include temperature, total precipitation, relative humidity, wind speed, hours of sunlight, and atmospheric pressure.The granularity of our dataset also allows us to include high-dimensional fixed effects to ({lambda }_{{{{rm{t}}}}}) and ({gamma }_{{{{rm{i}}}}}), to respectively control for any household-invariant and time-varying factors. Lastly, ({varepsilon }_{{{{rm{it}}}}}) is an idiosyncratic error term clustered two-way at the county and year-by-month level.DatasetThe dataset used in this analysis is compiled from two sources. First, household water usage is an unbalanced seven-year (from 1st January 2012 through 20th May 2019) panel of daily water usage from urban households located in 25 counties across seven provinces in China. The dataset contains 18,593,559 observations for 13,575 unique accounts. These seven provinces are mostly located in the heavily populated and economically active southern part of China, and the households are all located in apartment buildings (as opposed to standalone houses). The geographic distribution of observations is shown in Fig. S9. Installation of the meters was not a choice by the households. Rather, in most counties, developers are required to install ‘smart’ meters for newly-constructed buildings. Older buildings undergoing major renovations or retrofitting are also required to do the same. The dataset was obtained from a major company that specializes in installation of smart meters (Zhiheng Technology, http://www.gszh.cn), and represents all water meters installed by them as of May 2019.According to the company, once installed, these meters transmit daily water usage data to the installation company, using a combination of radio waves and cellular networks. Each daily observation contains the dwelling’s water meter ID, water usage, and location specified up to the neighborhood level (neighborhoods are the urban administrative equivalent of villages in rural areas). To improve data quality, we drop the entire year of observations for a household if: (1) there is no water usage for more than 60 days; (2) average daily water usage is less than 0.2 m3 or more than 0.8 m3 (at the 5th and 95th percentile of usage). Unlike household utilities such as electricity, it is more likely that observations of “no water usage” indicate that an apartment is unoccupied on that day.Following data cleaning, we collected water tariff reform information from a variety of online sources including government, water supply company, and news media websites. The advantages of a multi-county dataset are that it allows us to study different types of water tariff changes, and use treatment and comparison groups to identify policy impacts. A county is placed into the “treatment” group if they experienced the specific tariff reform that we are analyzing. Similarly, a county is included in the “control” group if they had already implemented the targeted tariff structure, and did not experience any price changes.Specifically, for the “flat-to-IBT” reform, the treatment group consists of households from eight counties that experienced a transition from flat water rates to IBT where unit price of water increases as usage increase. The control group consists of households in nine counties that maintained the same flat-rate tariff.For “within-tier” tariff reform, the treatment group includes households from four counties that experienced an increase in water price while remaining under the IBT structure while the control group consists of households from nine counties that maintained the same IBT structure.To fully maximize our sample, and exploit variation in policy implementation dates across the counties, there are several counties that were used respectively, as treatment and control groups in analyses of different tariff reforms (see Table S7 for full list). For instance, Wuyishan county is used as a treatment group in “flat-to-IBT” reform and as a control group in “within-tier”. In its first inclusion as a treatment county, we use data from 1st January 2012 through 20th May 2019, in which its water tariff changed from flat to IBT on 1st July 2016. In its second inclusion as a control county, we use data from 1 July 2016 through 20th May 2019, in which their IBT tariffs did not change.Similarly, there are some counties that experienced more than one tariff change. For example, Fuqing experienced a within-IBT change on 1st Apr 2015 and wastewater treatment fee increase on 1st Jan 2017. For the first inclusion, we used data from Fuqing for the period of 1st Jan 2012 – 31st Dec 2016, and for the second inclusion data from 1st Apr 2015 – 20th May 2019. By splitting the observation windows in this manner, we ensure that post-reform effects from one analysis do not contaminate the identification strategy of another. This approach maintains the internal validity of each reform evaluation while enabling efficient use of the dataset.As a first check of comparability between these treatment and comparison groups, we plot the distribution of households according to their water usage for each type of tariff change. We can see from Fig. S10 that the treatment and comparison groups show highly similar distributions across all both types of reforms. Balance tests for each type of tariff reform also show that the treatment and control counties are statistically identical on many socioeconomic characteristics (Table S8).The second source of data is daily weather information obtained from the 337 ground weather stations of the China Meteorological Data Service Center where we used inverse distance weighting to attribute weather for each county in the dataset57,58.As the water usage dataset is a non-random collection of counties located in southern provinces, we compared the socioeconomic characteristics of these counties with those of the larger provinces in which they are located. Table S9 presents the results of t-tests for mean values. Apart from urban residents’ per capita disposable income (which is statistically significant at the 10% level), there are no other statistically significant differences between the sample counties and their corresponding provinces, in gender and age structure, household size, economic development status, and rural residents’ disposable income. Therefore, we conclude that the sample in this paper is generally representative of southern China.Household survey on water tariffsTo complement the administrative dataset, we conducted household survey in China to assess residents’ understanding of water pricing systems and their behavioral responses to tariff reforms. The survey was administered via an online platform and collected 853 valid responses nationwide. Respondents were screened to include only individuals responsible for paying their household water bills, and participation was voluntary and anonymous. The questionnaire covered household water use, billing practices, and perceptions of tariff fairness, followed by randomized scenarios describing either a shift from flat-to-IBT tariffs or within-IBT uniform price increases. Respondents calculated expected changes in their monthly bills under each scenario, allowing us to measure comprehension of tariff structures. The survey concluded with demographic questions on age, gender, location, education, occupation, and income. Responses were pre-tested for clarity, and the final sample reflects a broad cross-section of urban and peri-urban households across China. The survey protocol was reviewed by the Institutional Review Board of Renmin University of China (L20250107), and determined to be exempt from full review under the category of minimal-risk, anonymous human-subjects research. No personally identifiable information was collected, and all participants provided informed consent prior to participation.Reporting summaryFurther information on research design is available in the Nature Portfolio Reporting Summary linked to this article.

    Data availability

    The data that support the findings of this study are available in Zenodo with the identifier: https://doi.org/10.5281/zenodo.1718029259.
    Code availability

    All data are processed and analyzed in Stata (16MP) and the code is available in Zenodo with the identifier: https://doi.org/10.5281/zenodo.1718029259.
    ReferencesGCEW. The Economics of Water: Valuing the Hydrological Cycle as a Global Common Good. (2024).United Nations. The United Nations World Water Development Report 2024. (United Nations, 2024).Savelli, E., Mazzoleni, M., Di Baldassarre, G., Cloke, H. & Rusca, M. Urban water crises driven by elites’ unsustainable consumption. Nat. Sustain. 6, 929–940 (2023).Article 

    Google Scholar 
    Andres, L. A. et al. Troubled Tariffs: Revisiting Water Pricing for Affordable and Sustainable Water Services. (2021).Alamanos, A., Sfyris, S., Fafoutis, C. & Mylopoulos, N. Urban water demand assessment for sustainable water resources management, under climate change and socioeconomic changes. Water Supply 20, 679–687 (2020).Article 
    CAS 

    Google Scholar 
    Li, M. et al. Exploring China’s water scarcity incorporating surface water quality and multiple existing solutions. Environ. Res. 246, 118191 (2024).Article 
    CAS 
    PubMed 

    Google Scholar 
    NBS. China Statistical Yearbook. (2023).Qin, P., Chen, S., Tan-Soo, J.-S. & Zhang, X.-B. Urban household water usage in adaptation to climate change: Evidence from China. Environ. Sci. Policy 136, 486–496 (2022).Article 

    Google Scholar 
    OECD. OECD Environmental Performance Reviews: Chile 2024. (2024).García-Valiñas, M. Á, Martínez-Espiñeira, R. & Suárez-Varela, M. M. Price and consumption misperception profiles: The role of information in the residential water sector. Environ. Resour. Econ. 80, 821–857 (2021).Article 

    Google Scholar 
    Brent, D. A. & Ward, M. B. Price perceptions in water demand. J. Environ. Econ. Manag. 98, 102266 (2019).Article 

    Google Scholar 
    Wang, X., Lee, J., Yan, J. & Thompson, G. D. Testing the behavior of rationally inattentive consumers in a residential water market. J. Environ. Econ. Manag. 92, 344–359 (2018).Article 

    Google Scholar 
    Billings, R. B. & Jones, C. V. Forecasting Urban Water Demand. (American Water Works Association, 2008).Heiss, F., Ornaghi, C. & Tonin, M. Inattention matters: An analysis of consumers’ inaction in choosing a water tariff. J. Eur. Economic Assoc. 21, 1686–1719 (2023).Article 

    Google Scholar 
    Carver, P. H. & Boland, J. J. Short-and long-run effects of price on municipal water use. Water Resour. Res. 16, 609–616 (1980).Article 
    ADS 

    Google Scholar 
    Asci, S., Borisova, T. & Dukes, M. Are price strategies effective in managing demand of high residential water users? Appl. Econ. 49, 66–77 (2017).Article 

    Google Scholar 
    Li, L. & Jeuland, M. Household water savings and response to dynamic incentives under nonlinear pricing. J. Environ. Econ. Manag. 119, 102811 (2023).Article 

    Google Scholar 
    Baerenklau, K. A., Schwabe, K. A. & Dinar, A. The residential water demand effect of increasing block rate water budgets. Land Econ. 90, 683–699 (2014).Article 

    Google Scholar 
    Lee, J., Nemati, M., Allaire, M. & Dinar, A. The impact of pricing structure change on residential water consumption: A long-term analysis of water utilities in California. Water Resour. Econ. 46, 100240 (2024).Article 

    Google Scholar 
    Nataraj, S. & Hanemann, W. M. Does marginal price matter? A regression discontinuity approach to estimating water demand. J. Environ. Econ. Manag. 61, 198–212 (2011).Article 

    Google Scholar 
    Wichman, C. J., Taylor, L. O. & Von Haefen, R. H. Conservation policies: Who responds to price and who responds to prescription? J. Environ. Econ. Manag. 79, 114–134 (2016).Article 

    Google Scholar 
    Wietelman, D. C., Wichman, C. J. & Brent, D. A. Conservation and distributional consequences of pricing scarce water during droughts. (Working Paper, 2024).El-Khattabi, A. R. et al. Heterogeneous responses to price: Evidence from residential water consumers. J. Environ. Econ. Manag. 107, 102430 (2021).Article 

    Google Scholar 
    Wichman, C. J. Perceived price in residential water demand: Evidence from a natural experiment. J. Econ. Behav. Organ. 107, 308–323 (2014).Article 

    Google Scholar 
    Schlosser, C. A. et al. The future of global water stress: An integrated assessment. Earth’s Future 2, 341–361 (2014).Article 
    ADS 

    Google Scholar 
    Mekonnen, M. M. & Hoekstra, A. Y. Four billion people facing severe water scarcity. Sci. Adv. 2, e1500323 (2016).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Shiferaw, B., Reddy, V. R. & Wani, S. P. Watershed externalities, shifting cropping patterns and groundwater depletion in Indian semi-arid villages: the effect of alternative water pricing policies. Ecol. Econ. 67, 327–340 (2008).Article 

    Google Scholar 
    Rogers, P., De Silva, R. & Bhatia, R. Water is an economic good: How to use prices to promote equity, efficiency, and sustainability. Water policy 4, 1–17 (2002).Article 

    Google Scholar 
    Mishra, B. K., Kumar, P., Saraswat, C., Chakraborty, S. & Gautam, A. Water security in a changing environment: Concept, challenges and solutions. Water 13, 490 (2021).Article 

    Google Scholar 
    Massarutto, A. Servant of too many masters: Residential water pricing and the challenge of sustainability. Uti. Policy 63, 101018 (2020).Article 

    Google Scholar 
    Zhang, B., Fang, K. & Baerenklau, K. A. Have C hinese water pricing reforms reduced urban residential water demand? Water Resour. Res. 53, 5057–5069 (2017).Article 
    ADS 

    Google Scholar 
    Han, S., Zhou, J., Liu, Z., Zhang, L. & Huang, X. Influence of community factors on water saving in a mega city after implementing the progressive price schemes. Water 13, 1097 (2021).Article 

    Google Scholar 
    Saez, E., Slemrod, J. & Giertz, S. H. The elasticity of taxable income with respect to marginal tax rates: A critical review. J. economic Lit. 50, 3–50 (2012).Article 

    Google Scholar 
    Olmstead, S. M., Hanemann, W. M. & Stavins, R. N. Water demand under alternative price structures. J. Environ. Econ. Manag. 54, 181–198 (2007).Article 

    Google Scholar 
    Callaway, B. & Sant’Anna, P. H. Difference-in-differences with multiple time periods. J. Econ. 225, 200–230 (2021).Article 
    MathSciNet 

    Google Scholar 
    Baker, A. C., Larcker, D. F. & Wang, C. C. How much should we trust staggered difference-in-differences estimates? J. Financial Econ. 144, 370–395 (2022).Article 

    Google Scholar 
    Gardner, J. Two-stage differences in differences. arXiv preprint arXiv:2207.05943 (2022).Chapagain, K. et al. Urban water security: A comparative assessment and policy analysis of five cities in diverse developing countries of Asia. Environ. Dev. 43, 100713 (2022).Article 

    Google Scholar 
    UNICEF & WHO. Progress on household drinking water, sanitation and hygiene 2000–2022: special focus on gender. (New York, 2023).He, C. et al. Future global urban water scarcity and potential solutions. Nat. Commun. 12, 4667 (2021).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Andres, L. A. et al. Doing more with less: Smarter subsidies for water supply and sanitation. (2019).Cardenas, H. & Whittington, D. Magnitude and distribution of electricity and water subsidies for households with private connections in Addis Ababa, Ethiopia. Water Econ. Policy 7, 2150004 (2021).Article 

    Google Scholar 
    Fuente, D. et al. Water and sanitation service delivery, pricing, and the poor: An empirical estimate of subsidy incidence in Nairobi, Kenya. Water Resour. Res. 52, 4845–4862 (2016).Article 
    ADS 

    Google Scholar 
    Brick, K., De Martino, S. & Visser, M. Behavioural nudges for water conservation in unequal settings: Experimental evidence from Cape Town. J. Environ. Econ. Manag. 121, 102852 (2023).Article 

    Google Scholar 
    Brick, S. Show, Don’t Tell: Emotion, Acquaintance and Moral Understanding Through Fiction. Br. J. Aesthet. 63, 501–522 (2023).Article 

    Google Scholar 
    Fielding, K. S. et al. An experimental test of voluntary strategies to promote urban water demand management. J. Environ. Manag. 114, 343–351 (2013).Article 

    Google Scholar 
    Wang, J. & Chermak, J. M. Is less always more? Conservation, efficiency and water education programs. Ecol. Econ. 184, 106994 (2021).Article 

    Google Scholar 
    Wichman, C. J. Information provision and consumer behavior: A natural experiment in billing frequency. J. Public Econ. 152, 13–33 (2017).Article 

    Google Scholar 
    Wheeler, S. A., Nauges, C. & Grafton, R. Q. Water pricing and markets: Principles, practices and proposals. Appl. Economic Perspect. Policy 47, 487–514 (2025).Article 

    Google Scholar 
    Kahan, J. P. Focus groups as a tool for policy analysis. Analyses Soc. Issues Public Policy 1, 129–146 (2001).Article 

    Google Scholar 
    Davies, A. R. Where do we go from here? Environmental focus groups and planning policy formation. Local Environ. 4, 295–316 (1999).Article 

    Google Scholar 
    Thaler, R. Mental accounting and consumer choice. Mark. Sci. 4, 199–214 (1985).Article 
    ADS 

    Google Scholar 
    Kookana, R. S., Drechsel, P., Jamwal, P. & Vanderzalm, J. Urbanisation and emerging economies: Issues and potential solutions for water and food security. Sci. Total Environ. 732, 139057 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Elenev, V., Quintero, L., Rebucci, A. & Simeonova, E. Staggered health policy adoption: Spillover effects and their implications. Management Science (2024).Huang, Y., Zou, C. & Ye, Z. Carbon emission reduction effect of innovative city pilot policies in China: based on the staggered difference-in-difference model. Environ. Sci. Pollut. Res. 30, 110812–110825 (2023).Article 

    Google Scholar 
    Karplus, V. J. & Wu, M. Dynamic responses of SO2 pollution to China’s environmental inspections. Proc. Natl Acad. Sci. 120, e2214262120 (2023).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Currie, J. & Neidell, M. Air pollution and infant health: what can we learn from California’s recent experience? Q. J. Econ. 120, 1003–1030 (2005).
    Google Scholar 
    Deschenes, O. & Greenstone, M. The economic impacts of climate change: evidence from agricultural output and random fluctuations in weather. Am. Econ. Rev. 97, 354–385 (2007).Article 

    Google Scholar 
    Tan-Soo, J.-S., Jun, L., Ping, Q. & Yu, Q. Data from: making water pricing work: how tariff familiarity sustains household conservation. Zenodo https://doi.org/10.5281/zenodo.17180292 (2025).Article 

    Google Scholar 
    Download referencesAcknowledgementsThe authors gratefully acknowledge financial support provided by the Asian Development Bank Institute for J.S.T.S., and Singapore Ministry of Education Tier 1 grant (A-8000736-00-00) for J.S.T.S. We are also grateful to the computational support provided by the Public Computing Cloud at Renmin University of China.Author informationAuthor notesThese authors contributed equally: Ping Qin, Jun Li.Authors and AffiliationsSchool of Applied Economics, Renmin University of China, Beijing, ChinaPing Qin & Jun LiNUS Business School, National University of Singapore, Singapore, SingaporeYu QinLee Kuan Yew School of Public Policy, National University of Singapore, Singapore, SingaporeJie-Sheng Tan-SooInstitute for Environment and Sustainability, National University of Singapore, Singapore, SingaporeJie-Sheng Tan-SooAuthorsPing QinView author publicationsSearch author on:PubMed Google ScholarJun LiView author publicationsSearch author on:PubMed Google ScholarYu QinView author publicationsSearch author on:PubMed Google ScholarJie-Sheng Tan-SooView author publicationsSearch author on:PubMed Google ScholarContributionsJ.S.T.S. conceived the study, and coordinated the overall research. J.S.T.S., P.Q., and J.L. designed the research. J.L. performed the analysis with support from J.S.T.S., and P.Q. P.Q. provided data. J.S.T.S., and P.Q. designed the survey, and J.L., and Y.Q. implemented the survey. J.S.T.S. wrote all versions of the manuscript. P.Q., J.L., and Y.Q. provided comments on the manuscript.Corresponding authorCorrespondence to
    Jie-Sheng Tan-Soo.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Peer review

    Peer review information
    Nature Communications thanks Johannes Reichl and the other, anonymous, reviewers for their contribution to the peer review of this work. A peer review file is available.

    Additional informationPublisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary informationSupplementary InformationReporting SummaryTransparent Peer Review fileRights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleQin, P., Li, J., Qin, Y. et al. Tariff familiarity sustains household water conservation.
    Nat Commun 16, 11235 (2025). https://doi.org/10.1038/s41467-025-66453-wDownload citationReceived: 03 March 2025Accepted: 05 November 2025Published: 18 December 2025Version of record: 18 December 2025DOI: https://doi.org/10.1038/s41467-025-66453-wShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Identification of tropical cyclone–related flash floods from hazard narratives using a large language model–based approach

    AbstractThis study employs a Large Language Model (LLM)–based approach to identify tropical cyclones (TCs)-related flash floods and understand their interrelationships using hazard narratives from the National Centers for Environmental Information Storm Events Database. A processing pipeline comprising two LLMs and a series of validation procedures was developed to extract the contributing factors of flash floods and classify them classified as either TC-related factors or external weather and environmental conditions. The analysis identified 6,470 flash floods across the contiguous United States related to 135 North Atlantic and East Pacific TCs between 2007 and 2024. The spatial proximity of flash floods to TC tracks varied substantially, suggesting multiple TC-related flood-generating mechanisms. The contributing factors to flash floods extracted from hazard narratives indicated that floods primarily driven by TC processes occurred closer to the TCs, whereas events influenced by external factors extended well beyond the typical TC rainfall range.

    Similar content being viewed by others

    Climate change exacerbates compound flooding from recent tropical cyclones

    Article
    Open access
    18 December 2024

    Basin-informed flood frequency analysis using deep learning exhibits consistent projected regional patterns over CONUS

    Article
    Open access
    13 April 2025

    The conterminous United States are projected to become more prone to flash floods in a high-end emissions scenario

    Article
    Open access
    06 April 2022

    IntroductionFloodings are among the world’s deadliest natural hazards, and have significant social, economic, and environmental impacts1. In the United States (U.S.), tropical cyclones (TCs) and their associated heavy rainfall are major causes of inland flooding2,3, and rainfall-induced flood deaths occurred in more TCs than in any other hazard4. Inland flooding can be broadly classified into two categories: flash flooding and riverine flooding. Flash floods are rapid flooding events occurring within six hours of a triggering event, typically heavy rainfall5. Owing to their rapid onset and complex hydrometeorological characteristics, flash floods pose severe threats to life, property, and infrastructure, and present significant challenges for forecasting, early warning, and flood risk management1,3,6. Recent deadly flash floods from Hurricane Helene in North Carolina (2024)7 and the remnants of Hurricane Hilary (2023) in the southwestern U.S.8 have drawn renewed attention to how TCs can interact with other weather systems and environmental conditions to produce flooding hundreds of kilometers away and even after the storm has dissipated. Research has found that most TC-related flood fatalities occur inland, where communities may not understand that they are at risk of TC-induced flooding3. Thus, there is a critical need for a comprehensive assessment of TC-related flash floods, including their spatial patterns and contributing factors. Such knowledge can strengthen public awareness, guide local preparedness efforts, and improve flood risk management.The climatology of TC-associated flooding in the U.S. has been extensively examined at local, regional, and continental scales2,9,10,11,12. Most of these studies identify floods as TC-associated when river discharge or flooding occurs within defined spatial and temporal windows of the storm track. A spatial window of 500 km from TC tracks is commonly used2,11, as previous studies have shown that this distance captures the majority of TC rainfall and effectively distinguishes TC rain from that produced by other weather systems13,14,15. Consequently, prior studies primarily focused on flash floods driven directly by the rainfall fields of active TCs whose locations were being tracked, overlooking events beyond these spatial ranges. However, previous studies have shown that TC- and remnant-induced environmental changes, particularly enhanced moisture, can interact with mid-latitude systems to produce flooding more than 1000 km away from the parent TC, including some of the most severe events on record16,17,18,19. Therefore, it is essential to incorporate these types of floods when assessing the full scope of TC-related hydrological and societal impacts.Previous studies seeking to identify the meteorological causes of flooding often require extensive datasets to characterize the complex interrelationships between storms and resulting flood events3,20. As demonstrated by prior work, these analyses typically rely on data from multiple mesoscale and synoptic-scale sources (e.g., weather charts, reanalysis data, and radar observations), along with visual inspection and expert judgment, to accurately attribute flood causation3,20. Consequently, such approaches can be time-intensive, making them impractical for large-scale or systematic assessments of TC-related flooding. In addition, most studies focus primarily on meteorological aspects, while other critical factors—such as soil moisture and topography21—are overlooked due to data limitations and the need for specialized expertise3.In contrast, narrative descriptions provided in hazard reports and individual studies capture the detailed features, context, and impacts of hazard events. These narratives, often prepared by domain experts, synthesize diverse qualitative information sources, making them invaluable for understanding the complex interrelationships among hazards. For example, one of the most comprehensive hazard databases in the U.S.—the National Oceanic and Atmospheric Administration (NOAA) National Centers for Environmental Information (NCEI) Storm Events Database22—includes narrative accounts describing the causes and impacts of recorded events. Recent advances in Natural Language Processing (NLP), particularly in Large Language Models (LLMs), now enable causal reasoning and the extraction of structured information from such large volumes of narrative data. Researchers have utilized NLP and LLMs in various tasks in the atmospheric science and hazard research, including identifying disasters interrelationships23, reasoning adverse weather conditions24,25, solving atmospheric science problems26, recognizing TC-related named entities27, and extracting disaster impacts28. Despite their promising capabilities, the potential and challenges of using LLMs and hazard narratives to advance the understanding of complex hazard interrelationships have yet to be fully explored.This study aims to employs an LLM-based approach to identify flash floods influenced by North Atlantic and East Pacific TCs across the contiguous United States (CONUS) from 2007 to 2024, using hazard narratives provided in the NCEI Storm Events Database. We developed a processing pipeline that integrates two LLMs with a series of validation procedures to extract factors causally contributing to flash floods from the narratives. The extracted factors were subsequently classified as either TC-related or external factors independent of TCs. We then examined the extracted results across three aspects. First, we provided an overview of TC-related flash flood events and their associated human and economic losses. Second, we measured the distance from flash floods to TC tracks to understand their spatial relationship. Third, we examined frequently contributing TC and external factors described in the narratives to better understand how TCs contributed to flash floods and their spatial characteristics. Together, these results provide an improved understanding of TC-related flash floods—where they occur and the conditions that contribute to them—which can support forecasting, mitigation, and public education. We also discuss the feasibility of using LLMs for large-scale analysis of text-based weather and hazard data, along with the study’s limitations and future research directions.ResultsOverview of TC-related flash floodsFrom 2007 to 2024, a total of 6470 flash floods in the CONUS were identified as related to North Atlantic and East Pacific TCs, representing 9.5% of all flash floods recorded during that time22. However, these TC-related flash floods caused 318 direct fatalities, accounting for 26.5% of the total 1197 flash flood deaths during the study period. About 2.3% of TC-related flash floods caused direct death, nearly double the 1.0% observed for all flash flood events6,22. In terms of direct property damage, TC-related flash floods resulted in a total of $64.1 billion (2024 USD), representing 73.2% of the total flash flood damage of $88.0 billion. This high proportion is largely driven by the extreme losses associated with Hurricane Harvey (2017)6,29. Overall, 38.7% of TC-related flash floods caused property damage, slightly higher than the 36.8% observed for all flash flood events.Figure 1 presents county-level TC-related flash flood frequency and damages from 2007 to 2024. A total of 1328 counties across 42 CONUS states plus Washington, D.C. experienced at least one TC-related flash flood event (Fig. 1a). In the east CONUS, the most affected areas were along the Atlantic and Gulf coasts, particularly the Carolinas and the Northeast. Florida showed a local minimum in flash flood frequency, consistent with prior studies indicating that much of its flooding is from slow-rise river floods rather than flash floods30. Although not a primary region directly impacted by TC landfalls, the Southwest still experiences frequent TC-related flood events. Even some northern areas, such as counties in Utah and Washington, have recorded such events10. The spatial distributions of these TC-related flood events are generally consistent with previous TC flooding studies2,10. In terms of damage, a total of 118 counties experienced at least one direct death caused by TC-related flash floods (Fig. 1b), and 806 counties experienced direct property damage (Fig. 1c). Both loss metrics reveal high damage clusters in Texas—primarily due to Hurricane Harvey (2017)29—and in western North Carolina, largely attributed to Hurricane Helene (2024)7. It is noteworthy that the damage does not show a sharp decline inland, as flash flood frequency does (Fig. 1a). Significant losses from TC-related flash floods occurred in inland regions that were otherwise less frequently impacted by TCs.Fig. 1: County-level distribution of TC-related flash floods during 2007–2024.a Event frequency. b Direct fatalities. c Directproperty damage (million 2024 USD). In all panels, darker shading indicates higher values, lighter shading indicates lower values,and and gray denotes counties without TC-related flash flooding or without reported fatalities or damage during the study period.Full size imageSpatial proximity of flash floods to TCsThis section examines the spatial relationship between storm location and flood onset. The spatial distribution of flash floods and the flood-inducing TC tracks are shown in Fig. 2. The locations of flash floods were represented by the midpoints between their reported start and end coordinates. These 6470 flash flood events were related to 135 TCs, including 90 North Atlantic TCs and 45 East Pacific TCs. Among these TCs, 65 TCs (44 East Pacific TCs and 21 North Atlantic TCs) did not make landfall over CONUS, and 31 TCs (26 East Pacific TCs and 5 North Atlantic TCs) never moved within 500 km of CONUS, such as Hurricane Joaquin (2015), which were related to severe flooding in South Carolina31. These results suggest that TCs can trigger flash floods far from the storm track. To quantify this, we measured the distance between each flash flood location and the nearest point on the corresponding storm track. For flash floods starting during active TC periods with valid track data, distances were calculated from the flood location to the TC track on the day the flood began. In addition, 17.9% of flash floods started after TC dissipation—when storm positions were no longer tracked. As a TC dissipates, its remnants often interact with and become increasingly influenced by large-scale weather systems, resulting in heavy rainfall10,19. For these flash floods, we measured distances from the flood locations to the nearest points along the entire TC track to assess proximity to the TC during its lifetime.Fig. 2: Locations of flash floods and related TC tracks during 2007–2024.a–f Flash flood locations and TC tracks for successiveperiods: a 2007–2009, b 2010–2012, c 2013–2015, d 2016–2018, e 2019–2021, and f 2022–2024. In all panels, blue pointsindicate reported flash flood locations, red lines represent TC tracks, and the gray buffer shows areas within 500 km of the contiguousUnited States.Full size imageTable 1 summarizes the flood-to-TC distances of flash floods occurring during and after the TC active lifecycle. For flash floods starting within the TC active lifecycle, the average distance from the flash flood to the same-day track segment is approximately 334.9 km, with half of the events located within 163.7 km. This aligns with previous findings that TC-related rainfall and flooding typically occur within a 500 km range2,32. About 22% of flash floods were located outside the 500 km of the storm track on the same day. The flash floods that began after TC dissipation occurred at significantly greater distances from the storm track, likely due to remnant lows being advected north or northwest and continuing to generate rainfall even after the TC was officially classified as dissipated19,33. The average distance was 896.8 km, with over 75% of these events occurring beyond 500 km from the entire TC track. Notably, the distance distributions for both groups exhibit large standard deviations, with maximum distances extending up to 2000 km.Table 1 Descriptive summary of the distance (km) from the flash flood events to the storm tracksFull size tableThe proximity of flash flood events to the storm track also shows geographic variation (Fig. 3). Figure 3a, b show the flood-to-TC distance occurring during the active TC lifecycle. For most counties in the coastal states along the Gulf and East Coasts, as well as along the inland trajectory from Texas northeastward to Ohio, the average distance is generally less than 500 km. This spatial pattern corresponds to the spatial pattern of landfalling North Atlantic TC tracks, which frequently generate flooding in close proximity to the storm path2. Small clusters of counties over the coast, along the Appalachian region and in the Midwest initiated flash flood events when the storm was, on average, positioned to the south, with over half of these events occurring at distances exceeding 500 km (Fig. 3b). The southwestern CONUS, which is primarily impacted by East Pacific TCs, experienced flash floods that tended to start when the TC was located farther away. Figure 4c, d illustrate the distance of the flash flood occurring after the active TC lifecycle. Post-TC dissipation events were concentrated in the Northeast and southwestern CONUS, with additional scattered clusters across the Midwest and Midsouth (Fig. 3d). On average, these events occurred at greater distances from the storm track, with numerous counties—especially in the Northeast, central, and western regions—recording events at distances exceeding 1000 km (Fig. 3c).Fig. 3: County-level spatial proximity of flash floods to related TCs.a Mean distance from flash floods beginning during the TCactive phase to the same-day TC track. b Percentage of flash floods occurring more than 500 km from the same-day TC track. c Mean distance from flash floods beginning after the TC active phase to the full TC track. d Percentage of flash floods occurringmore than 500 km from the full TC track. In all panels, darker shading indicates higher values, lighter shading indicates lower values,and gray denotes counties without events of this type.Full size imageIn summary, the spatial proximity of flash flood events to TCs exhibits great variability, with many occurring outside the typical spatial and temporal range of active TCs. The relatively long distance suggests that a TC-related floods can develop either prior to the arrival of a TC as a result of predecessor rainfall16,17,18, after the TC has passed or dissipated10,33, or from a TC that never directly approached the affected location31. These results also suggest that numerous floods are triggered not solely by TC rainfall, but also by complex interactions between TCs and external environmental factors.Contributing factors to TC-related flash floodsThis section examines how the TCs and external factors contribute to flash floods and their spatial characteristics. The 6470 flash floods were associated with 978 storm episodes, as defined in the Storm Events Database, which groups events linked to the same storm system34. Each episode includes a narrative summarizing the synoptic conditions and storm evolution. Further details are provided in the Methods section. The TC and external factors extracted from the narratives were normalized in categories and Fig. 4a shows the categories that are present in > 5% of 978 TC-related flood episodes and their interconnections.Fig. 4: Frequent TC and external contributing factors and their combinations.a Network graph of TC and external contributing factors present in more than 5% of episodes. Green nodes represent TC factors and orange nodes represent external factors; node size reflects episode frequency, and links indicate co-occurrence within the same episode. b Frequent combinations of TC and external factors occurring in more than 5% of episodes. Green bars indicate TC factors-only combinations, and orange bars indicate combinations involving both TC and external factors.Full size imageTC factors mentioned in these narratives predominantly fall into four categories: The name of the cyclone, remnants, TC rain, and moisture (Fig. 4a). TC names and remnants denote the corresponding storm systems. Almost all the episodes (~94.8%) refer to the storm by its name (e.g., Hurricane Ian) as expected, while about 46.4% of TC-related flood episodes explicitly indicate it’s the remnants of the storm that contribute to the flash floods. TC rain (~51.6% of the episodes) and TC-associated moisture (~40.5% of the episodes) were the two primary features identified as contributing TC factors to flash floods, together accounting for 84.8% of episodes. In the remaining episodes, TC roles were often implied through storm names or remnants only rather than explicitly stated.Approximately 45.3% of 978 flood episodes reported external factors contributing alongside TCs (Fig. 4a). Among these episodes, approximately 33.9% of episodes mentioned the influence of synoptic-scale systems, including fronts (~23.7%), upper-level enhancements (~10.7%), and low-pressure systems (~5.3%). These synoptic features can provide the lifting mechanisms for widespread heavy rainfall and flood-inducing storms3,16,20,35. Environmental instability and diurnal heating, which also contribute to lifting mechanisms for thunderstorm and heavy rainfall36,37, were reported in about 6.8% episodes. Approximately 6.6% of episodes reported the environmental moisture sources in addition to TC moisture, and about 5.2% of episodes reported the wind and flow features facilitating moisture transport, such as the low-level jet. These factors are consistent with previous research on ingredients that contribute to heavy rainfall that triggers flash floods37,38. In addition to atmospheric conditions, approximately 6.6% of episodes reported preexisting surface conditions (e.g., saturated soil from antecedent rainfall) or topography as contributing factors. Although less frequently mentioned, these factors are important considerations in flash flood forecasting21 and were associated with several high-impact events. For example, rainfall from Hurricane Matthew (2016) fell on already saturated soils caused by above-normal precipitation in September in North Carolina, resulting in catastrophic flash flooding39.We analyzed the co-occurrence frequency of these factors to assess how their interactions contribute to the flash floods. As TC names appeared in nearly all episodes, episodes were classified into two groups—active TCs and remnants—depending on whether the narratives explicitly referenced storm remnants. Figure 4b shows the most frequent factor combinations. First, active TCs and remnants exhibit different patterns: active TCs are more often linked to rainfall, while remnants contribute slightly more through moisture. Second, remnants and TC moisture commonly co-occurred with external systems such as fronts, upper-level features, and environmental instability. These findings are consistent with previous studies on TC-induced remote rainfall, which highlight that TCs supply abundant moisture that subsequently interacts with environmental lifting mechanisms to produce heavy precipitation40. As a TC dissipates and loses its organized structure, its remnants often interact with and become increasingly influenced by large-scale weather systems41.Because the combinations of factors reflect distinct TC–flood interrelationships, flash flood events were classified according to four distinct groups, namely active TC-only, remnants-only, active TC + external factors, and remnants + external factors, to examine their locations and spatial relationships to the TCs. Figure 5 shows box plots of flood–TC distances. The Kruskal–Wallis test with Dunn’s post-hoc analysis revealed significant differences (p < 0.01) in flood-to-TC distance among groups, except distance between the two groups involving external factors before TC dissipation (Fig. 5a). Figure 6 illustrates county-level flash flood frequency across the four groups.Fig. 5: Box plots of flash flood–to–TC distance for four groups.a Distances for flash flood events beginning during the TC active phase. b Distances for flash flood events beginning after TC dissipation.Full size imageFig. 6: County-level frequency of flash floods by event type.a Active TC-only events. b Remnants-only events. c Active TC + external factors events. d Remnants + external factors events. In all panels, darker shading indicates higher frequencies, lighter shading indicates lower frequencies, and gray denotes counties without that event type.Full size imageThe active TC–only group includes 2545 flash floods with no explicit indication of contributions from remnants or external factors. Approximately 96% of floods in this group occurred before TC dissipation. These events occurred close to the storm center at the time of flood onset, with an average distance of 175.2 km (St. Dev:204.1 km) (Fig. 5a). The extreme distance values (> 1182 km, 99th percentile) were associated with one North Atlantic TC (2020 Hurricane Cristobal) and four East Pacific TCs (2014 Hurricane Marie, 2015 Hurricane Dolores, 2018 Hurricane John, and 2023 Hurricane Hilary). This group of flash floods occurred mainly in the counties along the Gulf and East coasts (Fig. 6a). Two counties in the southwest also had a high frequency of active TC-only flash floods, while this region has fewer TCs directly hitting this area. We found that active TC–only events in these southwestern counties typically referenced the TC name or its moisture contribution, without explicit mention of external factors (e.g., Hurricane Hilary, 2023).The remnants–only group includes 1186 flash flood events associated with TC remnants, with no explicit indication of external contributing factors. Approximately 90% of floods in this group occurred before TC dissipation. These events were located further from the track with an average distance of 219.9 km (St. Dev:234.6 km) (Fig. 5a). The extreme distance values of this group (>1099 km, 99th percentile) were associated with two North Atlantic TCs (2015 Tropical Storm Bill and 2024 Hurricane Debby) and two East Pacific TC (2015 Hurricane Linda and 2022 Hurricane Kay). The remnants-only events primarily occurred along the Northeast coast, with scattered clusters in the mid-South, inland Texas counties, and the Southwest (Fig. 6b).The active TC + external factors group includes 1098 flash flood events with no clear mention of remnants but with reference to one or more external factors. Approximately 91% of floods in this group occurred before TC dissipated. As expected, the flash floods in this group occurred farther from the storm, averaging 608.2 km (SD: 533.4 km) from the storm track on the same day, with some events occurring as far as 1000–2000 km away (Fig. 5a). The Carolinas, as well as the southwest, observed the major clusters of this type of event (Fig. 6c).Lastly, the remnants + external factors group includes 1641 flash flood events with mentions of contribution from both TC remnants and external factors. Only about 51.2% of events started within the TC active cycle, averaging 609.1 km (St.Dev: 489.9 km) from the storm track on the same day (Fig. 5a). The remaining events in this group occurred after TC dissipation and represented the majority of flash floods observed in the post-dissipation period. These post-dissipation events were located with an average distance of 1032.7 km (St.Dev: 559.6 km) away from the any part of the entire TC track (Fig. 5b). Remnants + external factors events had a broad pattern from South Carolina to Maine, in the Midwest and inland South, and in the Southwest from southern California to New Mexico (Fig. 6d).DiscussionUnderstanding TC contributions to flash floods is crucial; however, conventional approaches often overlook events occurring beyond typical spatial ranges influenced by TC–environment interactions, largely due to data and method limitations. This study developed an LLM-based framework to extract contributing factors to flash floods from hazard narratives, identify TC-related events, and analyze their interrelationships. The results identified 6470 flash floods across the CONUS that were associated with 135 North Atlantic and East Pacific TCs during 2007–2024. Flash floods exhibited substantial variability in both location and timing relative to the parent TC. By examining the flash floods contributing factors extracted from the narratives, we found that approximately half of the flood episodes involved contributions from both TC-related and external environmental conditions. These differing flood-generation mechanisms are associated with distinct spatial characteristics of the flash floods. Flash floods driven solely by active TCs or the remnants typically occurred near the storm track during the TC’s active phase. In contrast, the flash floods that were influenced by TCs or their remnants in combination with external weather systems and environmental conditions tend to develop beyond TC rain range and after TC dissipation. Furthermore, TC–flash flood relationships exhibit clear spatial variability. Areas along the Gulf and Atlantic Coasts were more frequently affected by floods associated with active TCs, whereas northern and inland regions, such as Northeast, central, and Southwest, experienced a higher occurrence of flash floods contributed by TC remnants and environmental factors.To our knowledge, this study is among the first to employ LLM-based text mining of large-scale weather datasets to examine complex multi-hazard interactions. It advances the understanding of TC-related flash flood climatology by integrating both direct and indirect TC-induced events identified from expert-curated narratives. The quantification of the distance between flood events and TCs suggests that an inland location may experience flooding either before the cyclone makes its closest approach or even when it is not directly in the cyclone’s path. In addition, several of these remote flood events were associated with weaker storms and occurred after TC dissipation. This information can be valuable for improving public awareness and preparedness regarding flash flood safety, particularly for the inland regions beyond the direct TC track3. The findings also highlight the critical role of TC moisture and remnants, in conjunction with environmental factors, in generating floods—particularly those occurring beyond the typical spatial extent of TC influence. With an increase in TC moisture convergence and precipitation rate is anticipated due to climate warming, inland flood risks are expected to rise, highlighting the need for improved TC flood risk management, early warning, and emergency managment42,43.In addition, this work evaluated the feasibility of the proposed LLM-based framework. As detailed in the Methods section, the model demonstrated strong performance in extracting causal contributing factors to flash floods and labeling them as TC-related or external. The knowledge derived from the extracted information was broadly consistent with findings from previous studies using independent datasets and quantitative analyses, further confirming the robustness of the results. This study underscores the potential of LLMs for analyzing text-based weather and hazard data. Given that much meteorological information—such as forecasts, warnings, and post-event reports—is conveyed in textual form, applying LLMs offers a novel and effective approach to extract and synthesize insights that conventional analytical methods often overlook.This study has several limitations. First, the extracted information depends on the quality of the hazard report narratives. Although the Storm Events Database is considered one of the best publicly available collections of hazard reports in the U.S., and its records undergo quality control by local offices as well as the NCEI, biases and limitations may still exist due to reporting practices, human judgment, and varying data sources44. Second, LLM methods may have ambiguity or uncertainty depending on how the narrative describes the scenario. For example, while diurnal and topographic effects are important for heavy rainfall and flooding, these factors may be implied only through references to flood timing (e.g., afternoon) or location (e.g., Blue Ridge) rather than explicitly stated, leading to potential omissions in the extracted outputs. Future studies should carefully design and refine prompting strategies to better capture such implicit information within narrative text. Third, we presented contributing factor combinations for the same episode; however, complex spatial and temporal relationships may exist, as an episode can last up to five days34. Fully understanding the mechanisms of rainfall and flooding will require further work incorporating additional data and modeling approaches. For example, analysis of rainstorm structures as detected by ground-based radars can establish regions covered by TC rainbands and provide precise measurements of flood event distances from the TC’s edge.MethodsDataThis study obtained the storm names, tracks, and timing of TCs that originated from the East Pacific and North Atlantic basins from IBTrACS (International Best Track Archive for Climate Stewardship)45. A total of 654 TCs were included in the analysis to examine their contribution to the flash floods over the COUNS during the 18-year period from 2007 to 2024. The flash flood event information was obtained from the NOAA NCEI Storm Events Database, which documents severe weather events and unusual meteorological phenomena in the U.S. since 195022. This dataset is regarded as the best available and official archive of storm events in the U.S. and has been widely used in flash flood and weather hazards studies3,6,30,44,46.Inside the Storm Events Database, the data are structured in the following way. First, each record in the Storm Event database represents an individual hazardous event and its detailed information, such as hazard type (e.g., tropical storm, flash flood, or tornado), timing, locations, and loss information34. This work focused on flash flood events. The Storm Events Database reports several flood types, including flash floods, floods, coastal floods, and storm surges. The NCEI distinguishes flash floods from other flood types in the database. According to guidelines provided by the NWS, flash floods are defined as “a life-threatening, rapid rise of water into a normally dry area, beginning within minutes to several hours of the causative event (e.g., intense rainfall, dam failure, ice jam)”34. For each flash flood, the start and end times of the events were reported in the local time zone, and the locations were reported by the impacted county and by the start and end locations of the affected area in latitude and longitude coordinates. Flash flood losses include direct fatalities and property damages reported in actual U.S. dollars.Second, each event is part of a storm episode, which represents an entire storm system and may encompass various types of weather events. All events within a single episode are considered as associated with the same synoptic meteorological system34. Each episode includes an episode narrative, prepared by the NWS Weather Forecast Office that warned on the event, which provides a detailed textual description of the synoptic meteorology associated with the episode. For flooding events, a summary related to the rainfall or other conditions causing the floods are included. In particular, for TC-related events, the names of the TCs are included in the episode narratives of all related individual events. Since narratives have been available for all episodes only since October 2006, we focused on the period from 2007 to 2024 to ensure the analysis covered 18 complete hurricane seasons.Information extraction from the narrativesOur method for extracting TC and external factors casually contributing to flash floods from episode narratives uses a pipeline consisting of four main components, as illustrated in Fig. 7. The first component performed episode selection from the Storm Event database using a keyword filtering. The second component used LLM prompting to extract information about casual TC-related and external contributing factors to flash floods, storing the result in a semi-structured format. The third component conducted validation checks of the two LLMs outputs with the original narratives. The fourth component post-processed the semi-structured LLM outputs by exploring the word frequency and normalizing all extracted information. Below we describe each of the four components in more depth.Fig. 7: Information extraction pipeline with four main modules.a Text selection: flash flood episodes selected from the Storm Events Database using keyword filtering and filters on hazard type, time, and location. b LLM prompting: two LLMs used to extract contributing factors from the narratives. c Validation: LLM outputs checked against original narratives to generate the verified results. d Post-processing: extracted information analyzed through word-frequency exploration and normalization.Full size imageSelecting flash flood episode narrativesWe first filtered the Storm Events Database by hazard type, time, and location, identifying 20,394 episodes of flash flood events that occurred across the CONUS between 2007 and 2024. We then used a list of 654 East Pacific and North Atlantic TC names (e.g., “Ian” or “Helene) to perform a keyword-based search of the episode narratives, identifying episodes whose narratives contained a TC name corresponding to the year in which the TC occurred. This process yielded 1061 episodes involving flash floods that mentioned a TC name, which were used as inputs for LLM prompting. The episode narratives varied in length, with an average of 141 (St. Dev: 212) tokens per episode.LLM PromptingIn the core component of our information extraction pipeline, we fed the selected narratives into a LLM along with a sequence of prompts designed to identify causal contributing factors from TCs and external sources that are independent of TCs to the flash flood events. We constructed few-shot prompts to extract three levels of information: (1) whether the TC contributed to the flash floods; (2) the specific factors contributing to the flash floods; and (3) the classification of each factor as either TC-related or external. TC-related factors were defined as features or environmental conditions directly associated with, or primarily influenced by, TC. Examples include rainfall and thunderstorms generated by TCs, as well as tropical moisture transported by TC circulation. In contrast, external factors refer to weather systems or environmental conditions unrelated to TCs. Common examples of external factors include fronts, troughs, upper-level enhancement, topographic effects, atmospheric instability, diurnal heating cycles, and surface conditions37. We also requested the export of intermediate steps to aid in decision-making regarding how identified TC and external factors led to the flash floods. In particular, factors such as rainfall or thunderstorms may be classified as either TC-related or intermediate; the following method was applied to make this determination: If no external factors were mentioned in producing the rainfall, the rainfall was considered primarily a TC factor; otherwise, if the rainfall resulted from the interaction of a TC and external factors (e.g., a cold front), it was classified as an intermediate step. The prompts we applied are provided in Supplementary Text S1. We instructed LLM to generate outputs in CSV format to facilitate downstream processing and analysis.The prompt engineering process utilized several LLMs, including GPT-4.147, DeepSeek-v3.2-exp48, and Gemini-2.0-Flash (01-21)49. The formal analysis utilized the OpenAI GPT-4.1 and DeepSeek-v3.2-exp chat streaming APIs as the LLMs. During the inference stage, the temperature parameter is fixed at 0.3 to promote more deterministic outputs; the nucleus sampling threshold (top-p) is set to 1.0, ensuring that the entire probability mass is considered; the maximum token limit is determined by the default configuration of the respective models, and the top-k parameter is not specified in the streaming API.Validation and evaluation of LLM outputsThe raw output generated by the LLM required validation and consistency checks prior to inclusion in the analysis. To ensure data reliability, we conducted a series of validation procedures and compiled a verified list of TC-related episodes, along with their contributing factors as defined in the narratives. First, we identified the episodes that were related to TCs. Non–TC-related episodes were defined as those in which the LLM outputs either did not report any contributing factors, as instructed, or reported only external factors without any TC-related factors. At the episode level, the agreement between the two LLMs—OpenAI GPT-4.1 and DeepSeek-v3.2-exp chat—was 99.4%, with only six episodes showing disagreement. We manually went through all 1061 episodes to verify the results and ultimately identified 978 TC-related episodes.We combined the extracted factors from both LLMs for the 978 TC-related episodes and validated them against the original narratives. The two models collectively extracted 9409 contributing factors. We compared the outputs of the two LLMs using similarity metrics, including a fuzzy token set ratio threshold of 65 and a Jaccard similarity score threshold of 0.3. The two models exhibited an overall agreement of approximately 87%, with around 13% of the extracted factors differing between them. To ensure accuracy, all TC-related episodes were manually reviewed to validate the extracted TC and external factors against the original narratives, with particular attention given to unmatched cases. This process was supplemented by both exact and fuzzy word-matching techniques to determine whether the extracted information was explicitly present in the narrative rather than inferred or hallucinated by the models. Extracted factors were evaluated using defined criteria to ensure accuracy. A factor was considered a true positive if it (1) appeared in the original narrative or as a reasonable lexical variant; (2) represented a weather system or environmental condition contributing to the flash flood; and (3) was correctly labeled as either “TC” or “External.” Because LLM-generated boundaries can vary, similar expressions (e.g., “frontal boundary,” “a weak frontal boundary,” or “front”) were all accepted when contextually valid. After validation, a finalized list of TC-related and external contributing factors was compiled for all TC-related flash flood events.To address this ambiguity in entity extraction, we adopted an adjusted “manual score” approach suggested by Dagdelen et al.50 to determine whether the extracted phrase accurately reflects the source information, accounting for equivalent or variant expressions. Evaluation scores, including precision, recall, and F1, were calculated at both the episode and factor levels. While the precision (Eq. (1)) and recall (Eq. (2)) assess the prevalence of false negatives and false positives, the F1 (Eq. (3)) balances the importance of precision and recall and is preferable to accuracy for class-imbalanced datasets.$${precision}=frac{{No}.{of; true; postive}}{{No}.{of; true; positive}+{No}.{of; false; positive}}$$
    (1)
    $${recall}=frac{{No}.{of; true; positive}}{{No}.{of; true; positive}+{No}.{of; false; negative}}$$
    (2)
    $$F1=frac{2times {precision}times {recall}}{{precision}+{recall}}$$
    (3)
    The evaluation results are presented in Table 2. For the task of identifying whether a flash flood episode is related to a TC, both LLMs achieved a F1 score of 99%, demonstrating strong performance. The false negatives involved predecessor rainfall events or ambiguous cases where a TC was mentioned in the broader synoptic environment without a clear causal link to the flooding. The false positives primarily arose when TC name matches referred to place names rather than storms (e.g., Florence County, Ida County) or the TC being mentioned for unrelated purposes.Table 2 Model manual validation scoreFull size tableBoth models achieved F1 scores of 85–90% for extracting and labeling contributing factors. TC-factors showed marginally lower precision, largely from false positives involving non-contributing elements such as strong winds, tornadoes, or repeated terms like “flooding.” False negatives mainly reflected missed factors, such as cases where moisture was not correctly recognized as TC-related. For external factors, false positives mainly stemmed from misclassifying TC-related elements like tropical moisture as external, or from reporting non-contributing features (e.g., high tides). The false negatives were caused by the missing key environmental contributors to TCs. Although this does not constitute the most rigorous evaluation of the method, the strong agreement between the two models and with the original narratives nevertheless demonstrates the feasibility of extracting weather and environmental conditions and causal relationships from text-based data.Extracted information post-processingAfter validating the extracted factors, we performed a word frequency analysis using all confirmed contributing factors as an exploratory step. Word frequency was calculated after stop word removal and lemmatization, and the results were in the form of word clouds (Supplementary Figs. S1 and S2). These contributing factors generally align with previous research on the TC, heavy rainfall, and flash floods3,37. To prepare the data for quantitative analysis, each validated factor was mapped to a predefined category using keyword matching, fuzzy matching, and rule-based methods, guided by terminology from the National Weather Service Weather Glossary5, Glossary of National Hurricane Center Terms51, American Meteorological Society Glossary of Meteorology52, and prior studies on tropical cyclone rainfall and flooding3,20,35,37,38. The categories and exemplary factors are shown in Table 3.Table 3 Factor categorizing termsFull size tableSpatial and statistical analysisUtilizing the extracted information, we identified 978 TC-related episodes encompassing 6470 flash flood events. We first summarized flash flood events by their impacted counties, direct fatalities, and direct property damage. The direct property damage reported in the Storm Event database was adjusted for inflation using the consumer price index metric obtained from the U.S. Bureau of Labor Statistics53. All damage values were scaled to bring monetary values from the year of the event to the final data year (2024).Second, we analyzed the spatial proximity of flash flood events relative to their corresponding TCs using a Geographic Information System (GIS). Flash flood locations were represented by midpoints derived from start and end coordinates of impacted areas reported in Storm Events. Flash flood start times were converted to UTC and compared with TC active periods from IBTrACS, defined as the interval between TC origination and dissipation or cessation of tracking. For the flash flood events that started within the TC active cycle, the flood-to-TC distance was calculated from the flash flood midpoint to the nearest point on the TC track segment on the same day. For events that occurred after the TC’s active period—when no valid TC track was recorded—we measured the distance to the nearest point along the full TC track. All spatial analyses were performed using Python and ArcPy (ArcGIS Pro v3.3, Environmental Systems Research Institute, Inc.). These distance metrics were mapped at the county level to provide a geographic overview of where flash floods tend to start relative to TC locations.Last, we investigated the complex interactions between the TC and flash floods by examining the contributed factors to flash floods described in the narratives. Frequent TC and external factors and their interconnections were visualized using a network graph. Flash flood events were further classified according to their contributing factors: initially by the presence or absence of external factors, and subsequently by whether the episode was associated with a TC remnant. This classification resulted in four groups, namely active TC-only events, remnants-only events, active TC + external factors events, and remnants + external factors events. The Kruskal–Wallis test and Dunn’s post-hoc tests were applied to determine whether the distance metrics varied significantly among the four groups. Our working hypothesis is that flash floods influenced by TC and external factors occur at greater distances from the storm center.

    Data availability

    All the datasets analyzed in this study can be obtained from the following open sources. Tropical cyclone best track data are available at https://doi.org/10.25921/82ty-9e16. Storm Event Database is provided by the National Centers for Environmental Information at www.ncdc.noaa.gov/stormevents/.
    ReferencesWorld Meteorological Organization. Flash Flood Guidance System: Response to one of the deadliest hazards. World Meteorological Organization Available at: https://wmo.int/media/magazine-article/flash-flood-guidance-system-response-one-of-deadliest-hazards (accessed 17 Jul 2025). (WMO, 2020).Villarini, G., Goska, R., Smith, J. A. & Vecchi, G. A. North Atlantic tropical cyclones and U.S. flooding. Bull. Am. Meteorol. Soc. 95, 1381–1388 (2014).Article 

    Google Scholar 
    Ashley, S. T. & Ashley, W. S. The storm morphology of deadly flooding events in the United States. Int. J. Climatol. 28, 493–503 (2008).Article 

    Google Scholar 
    Rappaport, E. N. Fatalities in the United States from Atlantic tropical cyclones: new data and interpretation. Bull. Am. Meteorol. Soc. 95, 341–346 (2014).Article 

    Google Scholar 
    National Weather Service (NWS). National Weather Service Instruction 10–1605, NWSPD 10–16: Storm Data Preparation. National Weather Service Available at: https://www.weather.gov (accessed 5 Jul 2025). (NWS, 2021).Ahmadalipour, A. & Moradkhani, H. A data-driven analysis of flash flood hazard, fatalities, and damages over the CONUS during 1996–2017. J. Hydrol. 578, 124106 (2019).Article 

    Google Scholar 
    National Hurricane Center (NHC). Tropical Cyclone Report: AL092024 (Helene). National Hurricane Center Available at: https://www.nhc.noaa.gov/data/tcr/AL092024_Helene.pdf (accessed 17 Jun 2025). (NHC, 2024).National Hurricane Center (NHC). Tropical Cyclone Report: EP092023 (Hilary). National Hurricane Center Available at: https://www.nhc.noaa.gov/data/tcr/EP092023_Hilary.pdf (accessed 17 Jun 2025). (NHC, 2023).Villarini, G., Smith, J. A., Baeck, M. L., Marchok, T. & Vecchi, G. A. Characterization of rainfall distribution and flooding associated with U.S. landfalling tropical cyclones: analyses of Hurricanes Frances, Ivan, and Jeanne (2004). J. Geophys. Res. 116, D23 (2011).
    Google Scholar 
    Barth, N. A., Villarini, G. & White, K. Contribution of eastern North Pacific tropical cyclones and their remnants on flooding in the western United States. Int. J. Climatol. 38, 5441–5446 (2018).Article 

    Google Scholar 
    Liu, M., Smith, J. A., Yang, L. & Vecchi, G. A. Tropical cyclone flooding in the Carolinas. J. Hydrometeorol. 23, 53–70 (2022).Article 

    Google Scholar 
    Gori, A., Lin, N. & Smith, J. Assessing compound flooding from landfalling tropical cyclones on the North Carolina coast. Water Resour. Res. 56, e2019WR026788 (2020).Article 

    Google Scholar 
    Zhou, Y. & Matyas, C. J. Spatial characteristics of storm-total rainfall swaths associated with tropical cyclones over the Eastern United States. Int. J. Climatol. 37, 557–569 (2017).Article 

    Google Scholar 
    Jiang, H., Liu, C. & Zipser, E. J. A TRMM-based tropical cyclone cloud and precipitation feature database. J. Appl. Meteorol. Climatol. 50, 1255–1274 (2011).Article 

    Google Scholar 
    Prat, O. P. & Nelson, B. R. Precipitation contribution of tropical cyclones in the southeastern United States from 1998 to 2009 using TRMM satellite data. J. Clim. 26, 1047–1062 (2013).Article 

    Google Scholar 
    Galarneau, T. J., Bosart, L. F. & Schumacher, R. S. Predecessor rain events ahead of tropical cyclones. Mon. Weather Rev. 138, 3272–3297 (2010).Article 

    Google Scholar 
    Moore, B. J., Bosart, L. F., Keyser, D. & Jurewicz, M. L. Synoptic-scale environments of predecessor rain events occurring east of the Rocky Mountains in association with Atlantic basin tropical cyclones. Mon. Weather Rev. 141, 1022–1047 (2013).Article 

    Google Scholar 
    Rowe, S. T. & Villarini, G. Flooding associated with predecessor rain events over the Midwest United States. Environ. Res. Lett. 8, 024007 (2013).Article 

    Google Scholar 
    Ritchie, E. A., Wood, K. M., Gutzler, D. S. & White, S. R. The influence of eastern Pacific tropical cyclone remnants on the southwestern United States. Mon. Weather Rev. 139, 192–210 (2011).Article 

    Google Scholar 
    Mullens, E. D. Meteorological cause and characteristics of widespread heavy precipitation in the Texas Gulf watershed 2003–2018. Int J Climatol 41, 3743–3760 (2021).Article 

    Google Scholar 
    Gourley, J. J. & Clark, R. A. III. Real-time flash flood forecasting. In Oxford Encyclopedia of Natural Hazard Science (Oxford Univ. Press, https://doi.org/10.1093/acrefore/9780199389407.013.298 (2018).National Center for Environmental Information (NCEI). Storm Events Database. NOAA NCEI (2024). Available at: https://www.ncdc.noaa.gov/stormevents (accessed 20 Mar 2025).Liu, B., Guo, H. & Wang, H. An AI-driven approach to extract interrelationships between disasters. Int. J. Disaster Risk Reduct. 121, 105417 (2025).Article 

    Google Scholar 
    Zafarmomen, N. & Samadi, V. Can large language models effectively reason about adverse weather conditions? Environ. Model. Softw 188, 106421 (2025).Article 

    Google Scholar 
    Walczak, S. & Dinh, L. A text mining analytic approach for distinguishing between disaster and non-disaster zones from tweets. Int. J. Disaster Risk Reduct 118, 105233 (2025).Article 

    Google Scholar 
    Li, C., Deng, W., Lu, M. & Yuan, B. AtmosSci-Bench: evaluating the recent advance of large language model for atmospheric science. Preprint at https://doi.org/10.48550/arXiv.2502.01159 (2025) (accessed 27 Mar 2025).Zhou, Y. & Liu, P. Assessing multi-hazards related to tropical cyclones through large language models and geospatial approaches. Environ. Res. Lett. 19, 124069 (2024).Article 

    Google Scholar 
    Li, N. et al. Using LLMs to build a database of climate extreme impacts. In Proc. 1st Workshop on Natural Language Processing Meets Climate Change (ClimateNLP 2024) 93–110 (Association for Computational Linguistics, Bangkok, Thailand, https://doi.org/10.18653/v1/2024.climatenlp-1.7 (NLP, 2024).National Hurricane Center (NHC). Tropical Cyclone Report: AL092017 (Harvey). National Hurricane Center Available at: https://www.nhc.noaa.gov/data/tcr/AL092017_Harvey.pdf (accessed 17 Jun 2025). (NHC, 2018).Dougherty, E. & Rasmussen, K. L. Climatology of flood-producing storms and their associated rainfall characteristics in the United States. Mon. Weather Rev. 147, 3861–3877 (2019).Article 

    Google Scholar 
    Marciano, C. G. & Lackmann, G. M. The South Carolina flood of October 2015: moisture transport analysis and the role of Hurricane Joaquin. J. Hydrometeorol 18, 2973–2990 (2017).Article 

    Google Scholar 
    Jiang, H., Ramirez, E. M. & Cecil, D. J. Convective and rainfall properties of tropical cyclone inner cores and rainbands from 11 years of TRMM data. Mon. Weather Rev. 141, 431–450 (2013).Article 

    Google Scholar 
    National Weather Service. Climatology of tropical cyclone remnants for northern Illinois & northwest Indiana. https://www.weather.gov/lot/tropical_climatology. (accessed 17 Jun 2025).National Weather Service (NWS). National Weather Service Instruction 10–1605, NWSPD 10–16: Storm Data Preparation. National Weather Service (2021). Available at: https://www.weather.gov (accessed 5 Jul 2025).Konrad, C. E. & Perry, L. B. Relationships between tropical cyclones and heavy rainfall in the Carolina region of the USA Int. J. Climatol. 30, 522–534 (2010).Article 

    Google Scholar 
    Bowman, K. P. & Fowler, M. D. The diurnal cycle of precipitation in tropical cyclones. J. Clim. 28, 5325–5334 (2015).Article 

    Google Scholar 
    Lamers, A. et al. Forecasting tropical cyclone rainfall and flooding hazards and impacts. Trop. Cyclone Res. Rev. 12, 100–112 (2023).
    Google Scholar 
    Doswell, C. A., Brooks, H. E. & Maddox, R. A. Flash flood forecasting: an ingredients-based methodology. Weather Forecast 11, 560–581 (1996).2.0.CO;2″ data-track-item_id=”10.1175/1520-0434(1996)011<0560:FFFAIB>2.0.CO;2″ data-track-value=”article reference” data-track-action=”article reference” href=”https://doi.org/10.1175%2F1520-0434%281996%29011%3C0560%3AFFFAIB%3E2.0.CO%3B2″ aria-label=”Article reference 38″ data-doi=”10.1175/1520-0434(1996)011<0560:FFFAIB>2.0.CO;2″>Article 

    Google Scholar 
    Davis, C. Five years later, five lessons learned from Matthew–North Carolina State Climate Office. https://climate.ncsu.edu/blog/2021/10/five-years-later-five-lessons-learned-from-matthew/ (2021) (Accessed on Oct 31 2025).Schumacher, R. S., Galarneau, T. J. & Bosart, L. F. Distant effects of a recurving tropical cyclone on rainfall in a midlatitude convective system: a high-impact predecessor rain event. Mon. Weather Rev 139, 650–667 (2011).Article 

    Google Scholar 
    Lai, E. Tropical cyclone rainfall and flood forecasting. Global Guide Tropical Cyclone Forecast. (World Meteorological Organization, 2017).Knutson, T. et al. Tropical cyclones and climate change assessment: part II: projected response to anthropogenic warming. Bull. Am. Meteorol. Soc. 101, E303–E322 (2020).Article 

    Google Scholar 
    Maxwell, J. T. et al. Recent increases in tropical cyclone precipitation extremes over the US east coast. Proc. Natl. Acad. Sci. USA 118, e2105636118 (2021).Article 
    CAS 

    Google Scholar 
    Marjerison, R. D., Walter, M. T., Sullivan, P. J. & Colucci, S. J. Does population affect the location of flash flood reports? J. Appl. Meteorol. Climatol 55, 1953–1963 (2016).Article 

    Google Scholar 
    Knapp, K. R., Diamond, H. J., Kossin, J. P., Kruk, M. C. & Schreck III, C. J. M. International Best Track Archive for Climate Stewardship (IBTrACS) Project, Version 4. North Atlantic version. NOAA National Centers for Environmental Information (2018). https://doi.org/10.25921/82ty-9e16 (accessed 1 Dec 2024).Taszarek, M. et al. Severe convective storms across Europe and the United States. Part I: climatology of lightning, large hail, severe wind, and tornadoes. J. Clim 33, 10239–10261 (2020).Article 

    Google Scholar 
    OpenAI et al. GPT-4 Technical Report. Preprint at https://doi.org/10.48550/arXiv.2303.08774 (2024) (accessed 20 Mar 2025).DeepSeek-AI et al. DeepSeek-V3 Technical Report. Preprint at https://doi.org/10.48550/arXiv.2412.19437 (2025) (accessed 18 Apr 2025).Google Cloud. Gemini 2.0 Flash | Generative AI on Vertex AI. https://cloud.google.com/vertex-ai/generative-ai/docs/models/gemini/2-0-flash. (accessed 20 Mar 2025).Dagdelen, J. et al. Structured information extraction from scientific text with large language models. Nat. Commun. 15, 1418 (2024).Article 
    CAS 

    Google Scholar 
    National Hurricane Center. Glossary of NHC terms. https://www.nhc.noaa.gov/aboutgloss.shtml (accessed 27 May 2025).American Meteorological Society. Glossary of Meteorology. https://www.ametsoc.org/ams/publications/glossary-of-meteorology/(accessed 17 July 2025).Bureau of Labor Statistics. Consumer price index. https://www.bls.gov/cpi/ (accessed 24 Jul 2025).Download referencesAcknowledgementsY.Z. acknowledges the support from the Embry-Riddle Aeronautical University Faculty Research Development Program and the College of Aviation Small Equipment Grant. H.L. acknowledges the support from the U.S. National Science Foundation (DMS-2424605).Author informationAuthors and AffiliationsMeteorology Program, Department of Applied Aviation Sciences, Embry-Riddle Aeronautical University, Daytona Beach, FL, USAYao ZhouDepartment of Geography, University of Florida, Gainesville, FL, USACorene MatyasLinkedIn Corporation, Mountain View, CA, USAPing LiuDepartment of Geography and Sustainable Development, University of Miami, Coral Gables, FL, USAHan LiAuthorsYao ZhouView author publicationsSearch author on:PubMed Google ScholarCorene MatyasView author publicationsSearch author on:PubMed Google ScholarPing LiuView author publicationsSearch author on:PubMed Google ScholarHan LiView author publicationsSearch author on:PubMed Google ScholarContributionsY.Z. conceived the idea, collected the data, and conducted the preliminary analysis. Y.Z. and P.L. developed the methodology. Y.Z. and C.M. investigated the preliminary results. Y.Z. wrote the original draft. Y.Z., C.M., P.L., and H.L. reviewed and edited the manuscript.Corresponding authorCorrespondence to
    Yao Zhou.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Additional informationPublisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary information44304_2025_156_MOESM1_ESMRights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleZhou, Y., Matyas, C., Liu, P. et al. Identification of tropical cyclone–related flash floods from hazard narratives using a large language model–based approach.
    npj Nat. Hazards 2, 104 (2025). https://doi.org/10.1038/s44304-025-00156-6Download citationReceived: 21 August 2025Accepted: 26 November 2025Published: 18 December 2025Version of record: 18 December 2025DOI: https://doi.org/10.1038/s44304-025-00156-6Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Accelerating sea level rise in Africa and its large marine ecosystems since the 1990s

    AbstractSea-level rise poses a significant threat to Africa’s vital coastal ecosystems and the livelihoods of its growing populations. Here we analyze 31 years of satellite altimetry data to quantify sea-level change across Africa’s Large Marine Ecosystems, vast ocean regions of high biological productivity. The rate of rise has accelerated markedly to 4.34 mm/yr since 2010, over four times the 1990s rate. This is primarily driven by two factors: an increase in ocean mass from melting ice sheets accounts for over 80% of the total rise, with the remainder reflecting the expansion of warming ocean water. Regional rates are fastest in the Red Sea and Guinea Current, while increased salinity suppresses the trend in the Mediterranean. 2023 was particularly severe, with record-high sea levels across nearly 40% of Africa’s surrounding ocean. This uneven rise intensifies risks for over 50 million coastal residents, underscoring the urgent need for region-specific adaptation.

    Similar content being viewed by others

    Exploring steric sea level variability in the Eastern Tropical Atlantic Ocean: a three-decade study (1993–2022)

    Article
    Open access
    03 September 2024

    Sea level since the Last Glacial Maximum from the Atlantic coast of Africa

    Article
    Open access
    10 February 2025

    Modern sea-level rise breaks 4,000-year stability in southeastern China

    Article
    Open access
    15 October 2025

    IntroductionSea-level rise (SLR) stands among the most visible and rapidly accelerating consequences of anthropogenic climate change, posing an escalating threat to coastal systems globally. Manifesting through chronic flooding, shoreline erosion, saltwater intrusion, and the degradation of critical ecosystems, SLR has far-reaching consequences for both biodiversity and human livelihoods1,2. Since 1993, satellite altimetry missions including TOPEX/Poseidon, the Jason series, and Sentinel-6, have documented a persistent rise in global mean sea level (GMSL), currently estimated at 3.4 ± 0.3 mm/yr with an acceleration of 0.12 ± 0.05 mm/yr2,3,4. This increase is attributed to thermal expansion, along with ice mass loss from polar regions, glacier retreat and terrestrial water storage changes5,6. In 2023, ocean heat content in the upper 100 m reached unprecedented levels, with Marine heatwaves affecting up to 40% of the global ocean that further intensified steric contributions to sea level and exacerbated coastal vulnerabilities7,8.While SLR is a global phenomenon, its impacts are profoundly uneven and nowhere is this inequity more pronounced than in Africa (40°S–40°N, 25°W–60°E). Despite contributing to less than 4% of global CO₂ emissions, the continent bears a disproportionate burden of climate-related risks that threaten socio-economic stability, ecological integrity, and cultural heritage9,10,11. Africa’s 40,000-km coastline supports over 250 million people (Fig. 1), many of whom reside in low-lying urban centers like Alexandria, Cape Town, Dar es Salaam, Douala, and Lagos12,13. These cities are increasingly exposed to recurrent flooding, coastal erosion, and saltwater intrusion14,15,16,17. Moreover, local processes such as land subsidence and aquifer over-extraction amplify the effects of global SLR. In Lagos, for instance, subsidence rates of 8–12 mm/yr driven by intensive groundwater withdrawal significantly magnify relative SLR18,19,20.Fig. 1: Coastal vulnerabilities and oceanographic context of Africa.Population distribution across Africa, overlaid with the tracks of major tropical cyclones that have severely impacted East and Southern Africa over the past seven years: Alvaro (April 2019), Chalane (December 2020), Eloise (January 2021), Guambe (February 2021), Jobo (April 2021), Gombe (March 2022), and Freddy (February–March 2023). Solid lines indicate cyclone paths, while red and black dashed lines represent the minimum and maximum seasonal positions of the Intertropical Convergence Zone (ITCZ), marking the extent of the African monsoon. The boundaries of Africa’s seven LMEs are delineated in bold, ordered from highest to lowest annual ecosystem service value: Agulhas Current (ACLME, NOAA ID: 30), Red Sea (REDLME, 33), Somali Coastal Current (SCCLME, 31), Benguela Current (BCLME, 29), Guinea Current (GCLME, 28), Canary Current (CCLME, 27), and Mediterranean (MEDLME, 26), based on NOAA’s classification (https://www.st.nmfs.noaa.gov/ecosystems/lme/). Africa’s heritage location sites are marked with purple dots.Full size imageThe fragility of Africa’s coasts is further exacerbated by the deteriorating condition of its marine ecosystems. Africa’s seven Large Marine Ecosystems (LMEs) the Mediterranean, Red Sea, Somali Coastal Current, Agulhas Current, Benguela Current, Guinea Current, and Canary Current are ecological and economic lifelines that support diverse marine life and critical ecosystem services. However, these LMEs are under mounting pressure from both climate and anthropogenic stressors, including unregulated development, dam-induced sediment trapping, and habitat degradation21,22. Natural buffers like mangroves, coral reefs, and wetlands are steadily eroded, diminishing coastal resilience. At the same time, many African countries lack the infrastructure, early warning systems, and financial resources needed for adaptation23,24.These vulnerabilities are particularly acute in Eastern Boundary Upwelling Systems (EBUS) such as the Benguela and Canary Currents, which are among the world’s most productive marine systems22. These regions play an essential role in fisheries and climate regulation but are increasingly threatened by rising ocean temperatures, vertical stratification, and declining nutrient fluxes, all of which compromise marine productivity21,22,25,26. Meanwhile, non-EBUS regions, including the Somali and Guinea Currents, remain poorly studied despite their ecological and socio-economic importance. This highlights a critical knowledge gap that undermines effective adaptation planning in these regions. Island nations like Cape Verde, Seychelles, and São Tomé face even greater existential threats. With limited land area and economies closely tied to the ocean, they are especially vulnerable to the climate-related impacts of coral bleaching27,28, intensified storm surges, and shoreline retreat. The first two impacts are largely driven by ocean warming, whereas the third impact is a product of human intervention in coastal processes and SLR. These island states serve as stark microcosms of broader continental challenges, highlighting the urgency of regional SLR research.Compounding these regional vulnerabilities are large-scale climate oscillations and extreme weather events. The West African Monsoon (WAM), governed by the seasonal movement of the Intertropical Convergence Zone (ITCZ), delivers intense rainfall to the coastal zone, particularly the Gulf of Guinea29,30. Since 2000, the WAM has intensified, with a 15–20% increase in extreme rainfall events (≥85 mm/day), often resulting in devastating floods31. In 2023, a strong El Niño event caused a northward shift of the ITCZ, bringing unusually heavy rainfall and displacing over a million people in Cameroon, Ghana, Nigeria, and neighboring countries32,33. These events further exacerbate the compound nature of SLR and extreme rainfall, which together undermine the resilience of coastal communities15,34,35.Tropical cyclones and their associated storm surges further intensify the threat. Events such as Cyclone Eloise (2021), Tropical Storm Ana (2022), Cyclone Freddy (2023), and Storm Daniel (2023) have collectively affected more than 4 million people across Mozambique, Malawi, and Libya, with devastating consequences3,36,37 (Fig. 1). These rising seas are growing in frequency and intensity, outpacing the adaptive capacities of vulnerable nations. Notably, the 2023 failure of Libya’s Derna Dam during Storm Daniel was triggered by exposed structural vulnerabilities despite precipitation levels being within the dams’ design limits38. The destruction of Libya’s Derna Dam serves as a sobering reminder of the need for integrated, forward-looking disaster risk strategies.Understanding and projecting regional sea-level trends requires disentangling their physical components. Total SLR results from both steric (density-driven) and manometric (mass-driven) changes, with steric SLR subdivided into thermosteric (temperature-induced) and halosteric (salinity-induced) contributions39. In the Eastern Tropical Atlantic, recent studies show thermosteric contributions of 0.56 ± 0.03 mm/yr and ocean mass contributions of 2.66 ± 0.50 mm/yr14. However, halosteric trends, particularly in Africa’s LMEs, remain poorly constrained, limiting the accuracy of regional projections and adaptation planning. Further cloaking regional assessments are remote climate drivers. Teleconnection patterns such as the El Niño–Southern Oscillation (ENSO), the Indian Ocean Dipole (IOD), the Tropical Atlantic SST Index (TASI), and the North Atlantic Oscillation (NAO) modulate sea-level anomalies by influencing ocean circulation, vertical mixing, and thermal stratification7,14. The 2023 El Niño event, for example, not only amplified regional sea-level anomaly (SLA) variability but also intensified mean sea levels40,41, underscoring the importance of incorporating these remote signals into regional assessments.Despite the mounting risks, Africa remains underrepresented in global sea-level research. Most studies focus on global aggregates or select regions, particularly EBUS, leaving large portions of the African coastline unstudied. This uneven distribution of knowledge perpetuates an adaptation gap, undermining resilience in some of the world’s most vulnerable regions42,43. This underrepresentation highlights the urgent need for targeted studies across the entire African coastline.Drawing on three decades of satellite altimetry data from 1993 to 2023, this study aims to: (a) conduct an in-depth analysis of the record-breaking year 2023; (b) quantify updated SLA trends and disentangle the sea-level budget into the respective contributions of steric components (thermosteric and halosteric) and ocean mass, while incorporating Glacial Isostatic Adjustment (GIA) corrections; and (c) evaluate the influence of interannual climate modes on regional sea-level variability across Africa and its LMEs. By expanding the analysis beyond EBUS to include underrepresented LMEs and island nations, this study provides the first holistic synthesis of relative SLR dynamics across the African continent. In doing so, it addresses a critical gap in sea-level science and offers a robust foundation for evidence-based coastal planning and climate resilience.ResultsRegional sea level rise and the exceptional 2023 event in AfricaLong-term acceleration in African sea level rise (1993–2023)Satellite altimetry shows Africa’s regional mean geocentric sea level (RMSL) rose by 10.25 cm from 1993 to 2023, driven by a linear trend of 3.31 ± 0.04 mm/yr and an acceleration of 0.11 ± 0.02 mm/yr² (GIA-corrected; see Fig. 2a). Vertical land motion, while influencing relative SLR, does not affect altimetry-derived geocentric trends. This acceleration is comparable to the global rate (0.12 ± 0.05 mm/yr²)3,44. Notably, 2023 recorded the highest annual mean sea level to date. This increase is statistically significant, with a trend difference of 3.39 mm/yr between the first and last decades, and a notable 1.41 mm/yr difference between the last two decades (2003–2012 vs. 2013–2023). The rate of SLRFig. 2: Regional sea level variability and anomalies across Africa (1993–2023).a Time series of regional monthly mean sea level for Africa derived from satellite altimetry. Annual means are shown by red markers, with a purple curve representing a second-order polynomial trend. b Annual mean sea level for 2023 expressed as anomalies relative to the 1993–2022 baseline. c Difference between 2023 and 2022 annual means. d Spatial distribution of the year with the peak annual sea level recorded during the altimetry period. In 2023, sea levels across much of Africa continued to rise, primarily driven by global warming and modulated by large-scale climate variability such as El Niño. The seven African LMEs are labeled as in Fig. 1.Full size imagehas sharply increased, with cumulative SLRs (i.e., total increases over each decade) of approximately 0.92 cm (1993–2002), 2.82 cm (2003–2012), and 4.60 cm (2013–2023). The rate rose from 0.96 ± 0.26 mm/yr in the first decade to 4.34 ± 0.18 mm/yr in the last, a 4.54 ± 1.26-fold increase (Fig. 2a). Over the past two decades, the average rate (3.67 ± 0.16 mm/yr) was 1.2 times the long-term trend, highlighting ongoing acceleration since the early 1990s (Supplementary Fig. 1).This acceleration reflects intensified contributions from thermal expansion due to ocean warming and increased individual mass contributions, including glacier melt, Greenland and Antarctic ice sheets, changes in land water storage, freshwater fluxes, and atmospheric water vapor content45. These regional changes coincide with a globally coherent climate shift that began in the early 2010s, as evidenced by synchronized transitions in oceanic and atmospheric parameters worldwide, further underscoring the systemic nature of the observed SLR44. This rise is further amplified by mass contributions from melting glaciers, Greenland and Antarctic ice sheets, and changes in land water storage46,47, as depicted in the trend curve of Fig. 2a. While the long-term trend is generally upward, it includes minor fluctuations—for example, a subtle 0.09 cm decrease between 2021 to 2022 (ranking 22nd among 31 annual changes), likely influenced by regional ocean circulation anomalies. This was followed by a rise of 1.97 cm from 2022 to 2023—the second-highest annual increase on record, surpassed only by the 1997–1998 El Niño-driven rise. This increase drove 2023 to the highest annual domain-averaged SLA of 10 cm, with peak monthly values in November (12.50 cm) and December (12.73 cm) ranking as the top two months over the 1993–2023 period (see Supplementary Fig. 2). Notably, the 2022–2023 increase significantly outpaced the global mean rise of 0.59 cm.Complementing the temporal evolution, the spatial patterns in 2023 reveal an alarming extent of regional SLAs. Spatially, more than 95% of the African ocean surface exhibited elevated sea levels in 2023 compared to the 1993–2022 climatology — here defined as the average sea level over that 30-year period (Fig. 2b)—with more than 75% showing an increase relative to 2022. In 2023, monthly SLAs across Africa and its seven LMEs significantly exceeded the 1993–2022 climatology, with differences of 4.22–8.53 cm for Africa and up to 16.24 cm for the Somali Coastal Current (December). Other LMEs, including the Guinea Current (up to 8.67 cm, June), Benguela Current (up to 8.36 cm, December), and Mediterranean Sea (up to 11.22 cm, May), showed anomalies of 2.82–11.22 cm (Supplementary Fig. 2). Notably, 1.5% of the region experienced rises exceeding 10 cm, particularly in dynamic ocean systems such as the Agulhas Current (Fig. 2c). These localized hotspots of SLR are consistent with areas of strong mesoscale activity and boundary currents. By 2023, 38.6% of the regional ocean, especially in the Eastern Tropical Atlantic, reached its highest annual sea level on record, while 72.1% recorded peak levels during or after 2020 (Fig. 2d). This clustering of record-high sea level years since 2020 confirms that recent intensification is not isolated but rather part of a sustained spatial trend affecting large portions of the basin. This widespread occurrence of record-high sea levels in recent years points to a sustained and intensifying trend. These patterns align with global sea level trends, which show a 10.5 cm rise over the same period, but the accelerated rates and extreme annual increases in the African region amplify risks for coastal communities. The widespread elevation, affecting 95.8% of the region, underscores heightened coastal flooding risks, particularly in the Somali and Guinea Currents (Fig. 2b–d). These areas face heightened threats of inundation, salinization of freshwater resources, and loss of critical ecosystems.Peak SLA years and the exceptional 2023 eventWe assessed annual mean and peak monthly SLA, as well as their residuals, after removing the 1993–2023 GMSL trend (3.4 ± 0.3 mm/yr), as shown in the residuals time series in Fig. 3a. The analysis focused on five peak years: 1995, 1997, 2010, 2019, and 2023. Residuals were calculated by subtracting a linear GMSL trend fitted to satellite altimetry data for the African region, isolating spatially variable anomalies driven by regional oceanographic or climate processes. Peak months were identified as the months with the highest SLA per year. Notably, the 2023 values were exceptional. The annual mean SLA reached 67.95 mm, which was 19.24 mm higher than in 2019, and the December peak month SLA reached 77.65 mm, significantly exceeding 2019’s 56.66 mm, as illustrated in Fig. 3b. After removing the GMSL trend, the 2023 residuals remained outstanding: annual mean residuals were 14.13 mm, the highest of all years, and peak month residuals reached 22.32 mm, surpassing 1997’s value by 24.9%. Statistical tests, including Bonferroni-corrected t-tests48 (p < 0.0125 for 1997, 2010, 2019), supported the presence of persistent, regionally coherent anomalies, likely reflecting steric (temperature- and salinity-driven) or circulation-driven variability. Fig. 3b summarizes these metrics across peak years, underscoring 2023’s exceptional ranking in all categories and highlighting 2023’s dominance across all four metrics. Consequently, 2023 stands apart from years like 1997, which exhibited large residuals (16.64 mm) linked to the 1997–1998 El Niño, and 2019 Atlantic Niño, which showed high SLA (56.66 mm) but smaller residuals (14.53 mm). Trend analysis further supports this conclusion: SLA trends accelerated over the three decades, providing context for the residuals in Fig. 3a. This acceleration, combined with 2023’s large residuals, suggests contributions from both global and regional drivers, potentially linked to basin-scale climate variability, such as the 2023–2024 El Niño/Atlantic Niño or changes in coastal circulation. However, further analysis is needed to confirm these mechanisms. In conclusion, 2023 represents the most extreme regional SLA event in the satellite record for Africa, characterized by unprecedented magnitude, persistence after global signal removal, and strong regional forcing atop an accelerating background of global SLR.Fig. 3: Residual sea level anomalies in Africa highlight the exceptional nature of 2023.a Time series of SLA residuals after removing the 1993–2023 GMSL trend, with pink dots marking peak maxima. b Comparison of SLA and residual magnitudes for selected peak years. The year 2023 exhibits the largest anomalies in both SLA and residuals across the satellite record, persisting even after global trend removal.Full size imageSea levels trends and its componentsRegional trends variabilityIn this section, we analyze the spatial and long-term trend of the total SLA from 1993 to 2023, based on a reanalysis dataset that integrates altimetric, thermosteric, halosteric, and manometric components. This comprehensive approach captures the combined influence of these factors on sea level variability across the African region and its LMEs. We also quantify the relative contributions of the total steric effect and its individual components to overall sea level variability. As illustrated in Fig. 4a, the SLA trend map reveals significant spatial patterns, with a statistically significant (p < 0.05) trend in the RMSL. The color gradient (from blue to yellow) reflects the global mean SLR of 3.4 ± 0.3 mm/yr over the same period shown as a gray contour, which is used as a reference threshold44,49 to distinguish areas experiencing below- or above-average SLR compared to the global mean. Although the regional mean rate is similar to the GMSL, this average mask substantial spatial heterogeneity. Fig. 4a shows that much of the African oceanic region experiences sea-level anomaly (SLA) trends above the GMSL, particularly across several LMEs. Notably, the Guinea Current (GCLME), Canary Current (CCLME), Red Sea (REDLME), Somali Coastal Current (SCCLME), and the northern parts of the Benguela Current (BCLME) and Agulhas Current (ACLME) exhibit SLR rates exceeding the GMSL. These elevated rates may be linked to intensified wind-driven circulation or enhanced tropical thermosteric expansion. In contrast, the Mediterranean LME (MEDLME) (see Fig. 1) shows trends below the global mean.Fig. 4: Spatial distribution of regional sea-level trends across Africa’s coastal LMEs from 1993 to 2023.a Total sea-level trend; b total steric; c thermosteric; d halosteric; and e manometric (SLA minus steric) trends. GIA corrections were applied to both SLA and manometric trends. At each grid point, mean and seasonal cycles were removed prior to trend estimation. All trends are statistically significant at the 95% confidence level except in areas marked with gray dots. The seven African LMEs are labeled as in Fig. 1.Full size imageThe trends in SLA across Africa’s seven LMEs, as shown in Fig. 4a, reflect a compound interplay between steric (thermosteric + halosteric) and manometric components, each contributing differently across regions. These components interact to create distinct spatial patterns in SLA, shaped by both local and large-scale oceanographic processes. In the equatorial and tropical LMEs-such as the GCLME, CCLME, SCCLME, and the REDLME-thermosteric expansion, driven by ocean warming, is the principal contributor to SLR. The steric component, which encompasses changes in seawater density due to temperature and salinity, accounts for 19.78% of the total SLA in the entire region, as depicted in Fig. 4b. Notably, the thermosteric signal alone contributes 27.90% of SLA and exceeds the total steric trend by 41.10% (see Fig. 4c), underscoring the dominant role of thermal expansion. This effect is especially pronounced along the equatorial Atlantic and Indian Ocean margins, where persistent high temperatures and seasonal upwelling fuel significant thermal expansion.In contrast, subtropical regions such as the MEDLME display different dynamics. In the CCLME, thermosteric trends closely mirror SLA patterns, indicating that temperature-driven expansion remains important, as shown in Fig. 4a and Fig. 5 However, in the MEDLME, SLA variability aligns more closely with halosteric changes—those driven by salinity-evident in Fig. 4d. The halosteric signal strongly influences sea level trends in the MEDLME, where increased evaporation and reduced freshwater input elevate salinity, making it the dominant steric driver. The halosteric component, on average, exerts a modest but negative influence on SLA across most LMEs, contributing −8.13% (Fig. 4d). Negative halosteric trends are most notable in the MEDLME and broader North BCLME at the Congo River runoffs, likely linked to regional climate and hydrological patterns. Positive halosteric trends are mostly observed in the CCLME and ACLME, with localized positive trends in the GCLME, particularly off the coasts of Cameroon and Liberia, where major rivers discharge freshwater into the ocean, as highlighted in Fig. 4d. These regions, along with parts of the REDLME, Agulhas retroflection zone, and ACLME, experience enhanced steric SLR due to reduced water column density from riverine and estuarine runoff. Such freshwater inputs and regional circulation changes can locally offset the broader negative halosteric trend, highlighting the importance of hydrological variability in shaping regional sea level patterns. Dominating the SLA signal is the manometric component, which accounts for 80.23% of the total trend, as illustrated in Fig. 4e. Manometric sea level changes refer to variations caused by the redistribution of ocean mass, primarily from land ice melt, changes in terrestrial water storage, and large-scale shifts in the global water cycle. These changes are measured as the difference between total sea level changes (from satellite altimetry) and steric changes (from in situ temperature and salinity observations). In Africa, manometric trends generally range between 1.8 and 2.9 mm/yr in most LMEs. The highest values are observed in the SCCLME and GCLME, while lower rates (0.5–2 mm/yr) occur in the southernmost ACLME, MEDLME, and REDLME (Fig. 4e).Fig. 5: Temporal evolution and decadal sea level components for Africa.a Time series and linear trends of de-seasoned monthly anomalies of total SLA (blue), steric (SSLA; green), thermosteric (TSLA; red), halosteric (HSLA; purple), and manometric (orange) components from 1993 to 2023. Trend values for each component are indicated. All time series are smoothed using a 13-month low-pass filter to highlight interannual-to-decadal variability. b Decadal sea-level trends and component contributions for the African region, with bars showing trends (mm yr⁻¹) and error bars indicating uncertainties.Full size imageFigure 5a provides a comprehensive view of SLR trends over Africa by illustrating the evolution of the total SLA and its contributing components. The SLA (depicted by the blue line) shows a steady and significant upward trend of 3.30 ± 0.04 mm/yr since the early 1990s, closely aligning with the GMSL trend of 3.4 mm/yr. As shown in Fig. 5b, this regional trend masks a pronounced acceleration, increasing from 0.96 ± 0.26 mm/yr in 1993–2002 to 4.34 ± 0.18 mm/yr in 2013–2023, a more than fourfold increase, driven by two primary processes: steric changes, due to variations in seawater density (temperature and salinity), and manometric changes, linked to shifts in ocean mass from cryospheric and hydrological processes. The manometric contribution (orange line, Fig. 5a), with a trend of 2.61 ± 0.04 mm/yr, accounts for 79% of the SLA rise, dominating the total SLR. This is in line with Bellas-Manley et al.50, who found that the rise is primarily driven by mass inputs from global ice melt, especially Antarctic ice loss, which contributes up to 1.2 mm/yr through gravitational fingerprints in southern LMEs and Greenland’s more uniform effect, alongside changes in land water storage51. Manometric contributions have accelerated significantly, rising by 3.28 ± 0.11 mm/yr above 1993–2002 (Fig. 5b) levels, underscoring their pivotal role in shaping regional sea-level patterns. This acceleration closely tracks the intensified glacier melt reported by Dussaillant et al.52 with five of the last six years (2019, 2020, 2022, 2023, and 2024) exceeding 430 Gt/yr, reinforcing the substantial contribution of glaciers to the SLR acceleration observed along African coasts. The post‑2019 period of exceptional melt coincides with the smaller, but still positive, increment in manometric acceleration from the second to the third decade, indicating that while the largest step‑change occurred earlier (1993–2002 to 2003–2012), sustained high rates of glacier and ice‑sheet loss in recent years have maintained and reinforced elevated mass‑driven SLR trends across African LMEs. GIA has a negligible impact (<0.1 mm/yr) on African coasts, far from past ice sheet collapse regions. In contrast, the steric contribution (green line, Fig. 5a) is minor, with a trend of 0.34 ± 0.02 mm/yr, reflecting changes in seawater density. This is subdivided into thermosteric (temperature-driven) and halosteric (salinity-driven) components. The thermosteric anomaly (red line) exhibits a positive trend of 0.49 ± 0.02 mm/yr, indicating that ocean warming and thermal expansion are significant within the steric component. Conversely, the halosteric anomaly (purple line) shows a negative trend of −0.15 ± 0.01 mm/yr, suggesting that increasing salinity slightly counteracts SLR by increasing seawater density. The higher thermosteric trend compared to the total steric trend indicates that the halosteric component modestly offsets warming effects. While steric changes drive interannual variability, the manometric component, significantly driven by glacier melt as quantified by Dussaillant et al.52 dominates the accelerated SLR affecting African coastal regions.Figure 5 clarify the relative contributions, highlighting the dominance of the manometric component, followed by the thermosteric contribution, with the halosteric effect remaining negative and minor. All components exhibit temporal variability, with short-term oscillations superimposed on long-term trends.As shown in Fig. 5b, the African oceanic domains exhibit an increased trend over the observed last two decades, with the most recent decade marking an unprecedented increase in sea level, a signature of the strongest acceleration recorded to date. This regime shift aligns with the peak glacier mass loss documented by Dussaillant et al.52 particularly the 2023 record of 540 ± 69 Gt, contributing 1.5 mm to GMSL rise, which exacerbates regional SLR impacts on African coasts through enhanced ocean mass inputs. The steric components, in particular, show notable interannual variability, likely linked to climate phenomena such as ENSO and shifts in the ITCZ, which influence temperature and salinity distributions in the tropical ocean. Africa LMEs Sea level trends.
    Atlantic Ocean LMEs
    The CCLME, a productive eastern boundary upwelling system situated along northwest Africa in the Atlantic Ocean, exhibits a SLA trend of 3.30 ± 0.05 mm/yr over the study period Fig. 6a). This rise is primarily driven by mass (manometric) contributions of 2.63 ± 0.05 mm/yr, accounting for approximately 80% of the total and largely attributable to intensified land ice melt39,45. Steric contributions (0.66 ± 0.04 mm/yr), comprising about 20% of the SLA, include modest thermosteric expansion (0.47 ± 0.06 mm/yr) and a smaller positive halosteric signal (0.19 ± 0.03 mm/yr), the latter reflecting a salinity decrease primarily driven by increased continental runoff and possibly changes in precipitation, which overwhelms the salinizing effect of evaporation53,54. As a productive upwelling zone, the CCLME experiences moderated upper-ocean warming due to the persistent influx of cooler, nutrient-rich subsurface waters, which temper thermosteric expansion25,55. This upwelling-related cooling becomes more apparent in the decadal evolution of sea level trends. Between 1993–2002 (Fig. 9, Supplementary Table 1), SLA rose at 3.12 ± 0.30 mm/yr, driven largely by strong thermosteric expansion (2.07 ± 0.32 mm/yr). SLA peaked at 4.61 ± 0.31 mm/yr in 2003–2012, coinciding with an exceptional thermosteric contribution of 4.08 ± 0.23 mm/yr. However, in the most recent decade (2013–2023), although the SLA further increased to 4.69 ± 0.18 mm/yr, the thermosteric contribution dropped sharply to 0.45 ± 0.16 mm/yr—a decline of 1.60 mm/yr from the 1993–2002 (Fig. 9a) level-consistent with enhanced upwelling reducing heat retention in the upper ocean55. In parallel, mass contributions nearly doubled from 1.39 ± 0.18 mm/yr in 1993–2002 (Figs. 9a) to 3.12 ± 0.10 mm/yr in 2013–2023 (Fig. 9c), underscoring the increasing influence of land ice melt. Halosteric trends also shifted markedly from negative (−0.35 ± 0.04 mm/yr) to positive values (1.12 ± 0.07 mm/yr), reflecting a transition from salinity increase (which raises density and contributes negatively to sea level) to salinity decrease (which lowers density and contributes positively). This shift likely reflects enhanced freshwater input53. This shift supported a steric contribution of approximately 33% in the latest decade, with halosteric effects playing a more prominent role.Fig. 6: Regional sea level budget components for the Atlantic LMEs.Temporal evolution and linear trends of de-seasoned monthly anomalies for a the Canary Current LME (CCLME), b the Guinea Current LME (GCLME), and c the Benguela Current LME (BCLME) from 1993–2023. Contributions shown are total Sea-Level Anomaly (SLA; blue line), steric (SSLA; green line), thermosteric (TSLA; red line), halosteric (HSLA; purple line), and manometric (orange line). All time series are smoothed with a 13-month low-pass filter.Full size image
    Bordering the West African coast, the GCLME spans approximately 2 million km² and is characterized by seasonal upwelling, high freshwater input, and low-salinity surface waters25,56. Between 1993 and 2023, this region experienced a SLA trend of 3.41 ± 0.05 mm/yr, slightly above the global mean ranking it second among all LMEs (Fig. 6b).
    This rise is predominantly driven by mass (manometric) contributions of 2.68 ± 0.03 mm/yr, representing about 79% of the total. Steric effects contribute 0.72 ± 0.04 mm/yr (21%) to the regional sea level trend in the GCLME. Of this, thermosteric expansion estimated at 1.07 ± 0.05 mm/yr reflects upper-ocean warming across the tropical Atlantic, amplified by seasonal upwelling that redistributes heat in the upper layers, while halosteric changes (−0.35 ± 0.02 mm/yr) indicate contraction driven by salinity increase (i.e., density increase causing volume reduction). These findings align with recent studies emphasizing the joint role of mass and steric components in driving SLR in river-dominated tropical LMEs57. A decadal breakdown reveals a marked acceleration: SLA increased from 2.00 ± 0.27 mm/yr in 1993–2002 (Fig. 9a) to 4.50 ± 0.19 mm/yr in 2013–2023 (Fig. 9c), a 2.25-fold rise. Initially, this was supported by modest contributions from mass (0.54 ± 0.11 mm/yr) and thermosteric (1.55 ± 0.30 mm/yr) components. During 2003–2012 (Fig. 9b), both peaked, with manometric input reaching 3.63 ± 0.11 mm/yr and thermosteric expansion at 1.41 ± 0.18 mm/yr. In the most recent decade, mass contributions stabilized at 2.88 ± 0.09 mm/yr, while thermosteric trends remained steady at 3.07 ± 0.18 mm/yr, reflecting sustained ocean warming. Concurrently, halosteric contributions became increasingly negative, reaching −1.45 ± 0.06 mm/yr in 2013–2023 (Fig. 9c) consistent with increased salinity, which raises water density and reduces the height of the water column, thus lowering sea level. This intensification, as the GCLME transitions toward the Benguela system, underscores the growing dominance of mass-related SLR, compounded by thermal expansion and freshwater influxes, which are both projected to intensify under ongoing climate change55,58.
    The BCLME (Fig. 6c), stretching along the coasts of Angola, Namibia, and South Africa, is a wind-driven upwelling system and one of the world’s four major eastern boundary current upwelling regions59. From 1993 to 2023, the region recorded SLA trend of 2.98 ± 0.05 mm/yr (Fig. 6c), slightly below the GMSL. This rise is dominated by manometric (mass) contributions (2.52 ± 0.05 mm/yr), accounting for 85% of the total increase. A significant portion of the mass-driven rise is attributed to gravitational redistribution from Antarctic ice melt, which disproportionately affects the South Atlantic, and to regional upwelling dynamics that modulate local sea level55,57. Steric contributions remain modest (0.34 ± 0.03 mm/yr, 15%) due to the upwelling of cooler, nutrient-rich waters that suppress thermosteric expansion55. Thermosteric expansion is measured at 0.46 ± 0.03 mm/yr, but this is further reduced by slight halosteric contraction (−0.12 ± 0.02 mm/yr), indicating a net reduction in sea level contribution, likely driven by increased salinity due to enhanced evaporation or reduced freshwater input53. This balance of forces reflects the region’s sensitivity to both large-scale cryospheric changes (such as Antarctic meltwater) and local oceanographic processes (such as persistent upwelling)55. Decadal variability reveals a strong acceleration in SLA, rising from 2.13 ± 0.22 mm/yr in 1993–2002 (Figs. 9a) to 4.90 ± 0.17 mm/yr in 2013–2023 (Fig. 9c) more than doubling over the period, with a 2.3-fold increase. Early in the record, SLA was mainly driven by thermosteric expansion (2.65 ± 0.20 mm/yr) amid negligible or even negative manometric input (-0.23 ± 0.13 mm/yr). By the most recent decade, mass contributions surged 13-fold to 3.19 ± 0.10 mm/yr, reflecting intensified Antarctic meltwater input and the increasing dominance of mass-driven SLR in the region. Thermosteric trends also rose to 2.98 ± 0.16 mm/yr but were offset by strong halosteric contraction (−1.27 ± 0.09 mm/yr), driven by increased salinity, thereby reducing the net steric contribution to 15% of total SLR.

    Indian Ocean LMEs
    The ACLME, a warm western boundary current along South Africa’s east coast, presents a SLA trend of 3.00 ± 0.09 mm/yr (Fig. 7a), ranking sixth among the seven African LMEs analyzed, slightly above the BCLME. This long-term SLA trend is primarily driven by manometric contributions, which account for 2.30 ± 0.05 mm/yr (~77%). These mass-driven changes are attributed to regional ocean mass redistribution, influenced by the intense eddy activity characteristic of the ACLME. Steric contributions account for the remaining ~23% (0.70 ± 0.07 mm/yr), but their influence is comparatively modest. Within the steric component, thermosteric expansion ranges from 0.45 ± 0.03 to 0.67 ± 0.08 mm/yr, reflecting upper-ocean warming. However, this warming effect is dampened by persistent eddy-induced cooling and vertical mixing, which reduce the net thermal expansion60. Halosteric effects are minimal to slightly negative, ranging from −0.10 ± 0.02 to 0.02 ± 0.02 mm/yr (Fig. 7a), indicating weak salinity-driven contributions that marginally oppose thermal expansion. These effects further limit the overall steric contribution to the SLA trend. The region’s dynamic oceanography is characterized by strong decadal variability, driven by shifts in eddy activity, mass inflow, and upper-ocean heat content. During the period 1993–2002 (Fig. 9a), SLA exhibited a negative trend of −1.72 ± 0.65 mm/yr, driven primarily by strong cooling eddies and a large negative manometric contribution (−3.10 ± 0.20 mm/yr), while the thermosteric term was modestly positive (+1.45 ± 0.55 mm/yr) and the halosteric term slightly negative (−0.07 ± 0.06 mm/yr). These modest steric offsets were insufficient to counter the mass loss. This negative phase reflects reduced ocean mass input, linked to weakened current strength and shifts in regional atmospheric forcing, consistent with broader climate variability along the southeastern African coast.Fig. 7: Regional sea level budget components for the Indian Ocean LMEs.Temporal evolution of sea level budget components for a the Agulhas Current LME (ACLME) and b the Somali Coastal Current LME (SCCLME) from 1993 to 2023. Contributions shown are total Sea-Level Anomaly (SLA; blue line), steric (SSLA; green line), thermosteric (TSLA; red line), halosteric (HSLA; purple line), and manometric (orange line). All time series are smoothed with a 13-month low-pass filter.Full size image
    In contrast, the period 2003–2012 saw a sharp rise in SLA, with trends reaching 2.81 ± 0.35 mm/yr (Fig. 9b). This increase was supported primarily by rising manometric contributions (2.81 ± 0.18 mm/yr), associated with intensified mass inflow into the region. This positive shift aligns with strengthened eddy activity and increased regional circulation, which enhanced ocean mass accumulation and contributed to upper-ocean warming despite persistent mixing. Between 2013 and 2023, SLA stabilized at 1.12 ± 0.39 mm/yr, with mass contributions (2.45 ± 0.13 mm/yr) continuing to dominate (Fig. 9c). However, thermosteric trends remained negative (−2.02 ± 0.32 mm/yr), suppressing steric input to just 23% of the total rise. This underscores the ACLME’s sensitivity to dynamic ocean processes, where local steric effects are frequently modulated by turbulent mixing and eddy variability.
    Along East Africa’s Indian Ocean coast, the SCCLME unveils a SLA trend of 3.29 ± 0.13 mm/yr, ranking third among African LMEs (Fig. 7b). This trend is primarily driven by manometric contributions, which dominate the signal at 2.87 ± 0.06 mm/yr, accounting for 87% of the total rise. These mass-driven changes are largely attributed to global ice mass loss and monsoon-enhanced circulation, underscoring the critical role of ocean mass in shaping regional sea-level dynamics61,62. While smaller in magnitude, the steric contribution also plays a role in the SLA trend, with a net steric contribution of 0.42 ± 0.04 mm/yr. This is primarily driven by thermosteric expansion, which contributes 0.50 ± 0.04 mm/yr, reflecting modest ocean warming in the region. In contrast, halosteric effects are negligible, with a slight negative trend of –0.08 ± 0.02 mm/yr, indicating that increased salinity (density) slightly offsets the thermosteric expansion. This interplay between steric and manometric components highlights the dominance of mass-driven processes in the SCCLME, while steric contributions remain secondary. The sea-level budget also reveals pronounced decadal variability, reflecting the dynamic interplay of physical processes over time. During the period 1993–2002 (Fig. 9a), SLA trends were negative at –1.90 ± 0.88 mm/yr, driven by declines in both manometric (–2.97 ± 0.24 mm/yr) and thermosteric (1.01 ± 0.79 mm/yr) contributions. This period of reduced SLR coincided with weak monsoonal forcing63, which likely suppressed ocean circulation and upper-ocean heat content, reducing both mass and steric contributions. In contrast, the period 2003–2012 marked a sharp acceleration in SLA, with trends peaking at 5.92 ± 0.44 mm/yr (Fig. 9b). This dramatic increase is likely to be driven by strengthened monsoonal winds64 and enhanced ocean warming, as reflected in significant gains in both manometric (3.93 ± 0.12 mm/yr) and thermosteric (2.28 ± 0.43 mm/yr) contributions. The intensified monsoonal forcing during this decade likely enhanced upper-ocean heat content and circulation, amplifying both steric and mass-driven SLR. The most recent decade, 2013–2023, shows a sustained SLA trend of 4.72 ± 0.67 mm/yr (Fig. 9c), slightly lower than the previous decade but still elevated. This trend remains primarily driven by manometric contributions (3.44 ± 0.15 mm/yr), which continue to dominate the sea-level budget with an 87% share.
    However, the thermosteric contribution decreased to 1.50 ± 0.25 mm/yr, reflecting a slight reduction in ocean warming compared to the previous decade. Despite this decline, the dominance of mass-driven processes persists, underscoring the critical role of global ice mass loss and monsoonal dynamics in shaping regional sea-level trends. Looking ahead, the anticipated succession of El Niño and La Niña phases is expected to intensify interannual variability in the SCCLME. These climate oscillations are likely to modulate monsoonal wind patterns, upwelling strength, and upper-ocean heat content, potentially amplifying both thermosteric and manometric contributions to regional SLR.

    Semi-Enclosed Basin LMEs
    The REDLME exhibits the highest SLA trend among African LMEs, measured at 3.91 ± 0.13 mm/yr (Fig. 8a). This pronounced trend is primarily driven by the manometric contribution (2.06 ± 0.08 mm/yr), which accounts for 53% of the total rise, while thermosteric expansion (1.84 ± 0.09 mm/yr) accounts for 47%. The confined bathymetry and limited water exchange of the Red Sea enhance heat retention, amplifying upper-ocean warming and contributing significantly to the observed thermosteric trends. In contrast, halosteric effects (0.01 ± 0.02 mm/yr) are negligible, indicating minimal salinity-driven density changes and limited impact on steric sea level variations in this region. Together, the steric contribution totals 1.85 ± 0.09 mm/yr, underscoring the combined influence of temperature and salinity on SLR in this region. The decadal variability of SLA in the Red Sea LME highlights the dynamic interplay between regional temperature, salinity, and mass-driven processes. During the period 1993–2002 (Fig. 9a), SLA exhibited a strongly negative trend of −7.14 ± 0.51 mm/yr, driven by significant declines in both thermosteric (−4.27 ± 0.27 mm/yr) and manometric (−3.25 ± 0.37 mm/yr) contributions. This period of negative trends reflects regional cooling and reduced mass inputs, likely linked to weakened trade winds65 and lower heat retention. In contrast, the period 2003–2012 (Fig. 9b) marked a dramatic reversal, with SLA peaking at 6.18 ± 0.59 mm/yr. This sharp increase was driven by thermosteric contributions (2.89 ± 0.60 mm/yr), reflecting intensified upper-ocean warming, and manometric contributions (2.66 ± 0.27 mm/yr), associated with wind-driven mass inflow. The warming trends during this decade highlight the Red Sea’s sensitivity to regional and global climate forcing, as its confined geography amplifies heat accumulation and limits heat dissipation. By the period 2013–2023 (Fig. 9c), SLA trends stabilized at 3.55 ± 0.43 mm/yr, with thermosteric expansion (3.23 ± 0.40 mm/yr) emerging as the dominant driver. During this time, manometric contributions dropped to 0.01 ± 0.42 mm/yr, indicating a reduced influence of mass-driven processes, such as wind-driven inflow or regional mass redistribution. The sustained thermosteric trends, however, underscore the persistent warming of the Red Sea’s water column, which continues to drive SLR in the absence of significant salinity-driven or mass-driven changes. The observed SLA trends in the Red Sea LME are closely tied to the region’s unique oceanographic conditions, limited water exchange, and sensitivity to atmospheric forcing. The strong connection between SLA and temperature trends in the water column highlights the critical role of heat retention in driving thermosteric expansion. While salinity trends remain minimal, their slight variability contributes to the overall steric balance. The decadal shifts in SLA further emphasize the influence of regional climate variability, including trade wind patterns and heat fluxes, on the Red Sea’s dynamic sea-level budget.Fig. 8: Regional sea level budget components for the Red Sea and Mediterranean Sea LMEs.Temporal evolution of sea level budget components for a the Red Sea LME (REDLME) and b the Mediterranean Sea LME (MEDLME) from 1993 to 2023. Contributions shown are total Sea-Level Anomaly (SLA; blue line), steric (SSLA; green line), thermosteric (TSLA; red line), halosteric (HSLA; purple line), and manometric (orange line). All time series are smoothed with a 13-month low-pass filter.Full size imageFig. 9: Decadal sea level budgets across African LMEs.Histograms of decadal sea-level budgets for a 1993–2002, b 2003–2012, and c 2013–2023 across Africa’s LMEs.Full size image
    The MEDLME, the northernmost of the African LMEs, exhibits the lowest SLA trend among them, with an average rise of 2.70 ± 0.06 mm/yr from 1993 to 2023 (Fig. 8b), representing the lowest trends across African LMEs during the analysis period (see Table 1). This reduced SLA trend is primarily attributed to strong negative halosteric effects (-1.03 ± 0.04 mm/yr) caused by reduced Nile River discharge and high evaporation rates, which increase salinity and density, thereby counteracting thermosteric expansion. In contrast, other African LMEs, such as the Red Sea (3.91 ± 0.13 mm/yr) and the Guinea Current (3.41 ± 0.05 mm/yr), exhibit higher SLA trends due to the dominance of mass-driven contributions and weaker salinity-driven suppression. The manometric contribution in the Mediterranean, which includes the effects of oceanic mass change and redistribution and changes in bottom pressure caused by ocean currents55, accounts for 1.78 ± 0.07 mm/yr, representing 66% of the total SLA trend. Groundwater extraction and tectonic activity may also contribute to the residual trend. Meanwhile, thermosteric expansion contributes 1.95 ± 0.04 mm/yr, but this is significantly offset by salinity-driven contraction, resulting in a net steric contribution of 0.92 ± 0.05 mm/yr (34%). Decadal trends reveal substantial variability in SLAs, reflecting the dynamic interplay of physical oceanographic processes. Between 1993 and 2002 (Fig. 9a), SLA rose at 3.94 ± 0.32 mm/yr, primarily driven by strong thermosteric expansion (4.50 ± 0.10 mm/yr) associated with upper-ocean warming66,67, while manometric contributions were minimal (−0.35 ± 0.26 mm/yr), underscoring the dominance of temperature-induced changes during this period. From 2003 to 2012 (Fig. 9b), SLA accelerated to 5.32 ± 0.42 mm/yr, coinciding with a significant rise in manometric contributions (3.18 ± 0.42 mm/yr) linked to ocean mass redistribution, bottom pressure changes, and wind-driven circulation68,69. Meanwhile, steric contributions weakened to 2.13 ± 0.22 mm/yr as emerging negative halosteric trends (−1.88 ± 0.15 mm/yr) began to offset thermosteric expansion. In the Mediterranean Sea, SLA deceleration was primarily driven by intensified salinity effects: negative halosteric trends, indicating increased salinity, reduced SLA by 0.82 mm/yr, thereby diminishing both thermosteric and manometric contributions66,70.Table 1 SLA and contributing trends from 1993 to 2023 (mm/yr)Full size table
    This highlights the Mediterranean as a key region for understanding salinity-driven suppression of SLR. During the most recent decade (2013–2023, Fig. 9c), SLA growth slowed to 2.71 ± 0.30 mm/yr, representing a 31% decrease from the 2003–2012 peak and 1.24 mm/yr below the early period’s rate. This decline reflects a shift in ocean dynamics, with steric contributions accounting for only 34% of the total SLA during this period. Negative steric trends (−0.93 ± 0.20 mm/yr), dominated by halosteric suppression (−2.86 ± 0.07 mm/yr), played a key role in this slowdown. While unusually large for a halosteric contribution, this value is supported by the data and may reflect intense regional salinification. Although manometric contributions increased to 3.64 ± 0.34 mm/yr, their effect was largely offset by the persistent influence of salinity. The Mediterranean’s low SLA trend is justified by its unique salinity-driven suppression of SLR, which distinguishes it from other African LMEs where mass-driven SLR typically dominates. For example, the Red Sea exhibits the highest SLA trend (3.91 ± 0.13 mm/yr, Fig. 8a) due to strong manometric contributions from ice melt and water mass redistribution, while the Guinea Current (3.41 ± 0.05 mm/yr, Fig. 6b) and Canary Current (3.30 ± 0.05 mm/yr, Fig. 6a) are similarly influenced by mass-driven processes. In contrast, the Mediterranean’s sensitivity to hydrological changes, including reduced riverine input, high evaporation, and regional climate variability, results in a distinct response characterized by salinity-driven contraction. The inclusion of manometric contributions, which encompass oceanic mass redistribution, bottom pressure changes, and local vertical land movements, helps disentangle the competing drivers of SLA in this region and reinforces the central role of salinity changes in modulating sea-level trends.
    Ocean-atmosphere interactions and sea level variability in African LMEsClimate modes drive large-scale patterns of sea level variability across ocean basins, as noted by Han et al.71. In the context of Africa’s seven LMEs, the influence of these modes varies between the open ocean and coastal regions, particularly along the EBUS of the Canary and Benguela Currents. SLAs along these eastern boundaries are predominantly influenced by remote and local wind forcing, which propagates through equatorial and coastal waveguides14,34. In contrast, SLAs in the interior of these LMEs are driven by open-ocean forcing that propagates westward from the eastern boundaries. Consequently, coastal sea level variability in African LMEs is expected to correlate strongly with climate modes that are prominent in the tropics and with atmospheric centers of action that modulate longshore winds and sea level pressure.The regression of SLA with respect to climate indices is shown in Fig. 10 (the corresponding correlation map is provided in Supplementary Fig. 3). Starting with the Atlantic Niño (ATL3) in Fig. 10a, a strong positive regression emerges along both Atlantic and equatorial domains, especially along the Gulf of Guinea and extending westward. This pattern reflects the role of ATL3 in modulating SLA through SST anomalies, which drive thermal expansion and mass redistribution in the ocean. The influence of ATL3 is most pronounced in the GCLME, BCLME, and SCCLME, but it diminishes both northward toward the Canary and southward toward the Benguela Current, indicating a more localized impact of ATL3 on SLA variability, largely confined to the tropics. Transitioning to the Tropical North Atlantic (TNA) index in Fig. 10b, the regression map reveals pronounced positive anomalies along the northern tropical Atlantic, with significant effects extending into the Canary Current LME and the Mediterranean Sea. This underscores the importance of SST anomalies in the Tropical North Atlantic, which drive SLA variability through atmospheric teleconnections and oceanic processes such as wind-driven circulation and heat fluxes. However, the influence of TNA fades in the southern LMEs, including GCLME, BCLME, ACLMEs, and SCCLME, highlighting the regional specificity of this index and its primary relevance to the northern Atlantic margin with no impact on the REDLME. The Tropical Atlantic SST Gradient Index (TASI), shown in Fig. 10c, captures the meridional SST gradient in the Atlantic and exhibits a dipolar pattern, with positive regression coefficients in the northern Atlantic while not significant as the TNA and negative coefficients in the southern Atlantic. This spatial structure emphasizes the role of TASI in modulating SLA through changes in the interhemispheric SST gradient, which in turn affects wind patterns and ocean circulation. The Canary and Benguela Current LMEs are particularly sensitive to TASI, as these regions are directly influenced by shifts in the Atlantic dipole mode, while the Mediterranean LME and other LMEs remain less affected. Moving eastward, Fig. 10d presents the regression of SLA on the Indian Ocean Dipole (DMI). Here, strong positive coefficients dominate the western Indian Ocean, especially along the Somali Current LME and extending into the Agulhas Current LME. This dipolar pattern reflects the DMI’s capacity to drive SLA variability through SST changes and associated atmospheric circulation, with the Somali Current LME being especially responsive to the western Indian Ocean’s warming and mass redistribution. The Atlantic LMEs, by contrast, show minimal response to DMI, underscoring the basin-specific nature of this influence. The Western Tropical Indian Ocean (WTIO) index, depicted in Fig. 10e, further reinforces the dominance of the western Indian Ocean, with a pronounced positive regression along the Somali and Agulhas Current LMEs. The influence of WTIO even extends into the Red Sea Current outlets through the Gulf of Aden, highlighting the significant role of SST anomalies in this region in driving SLA variability through thermal expansion and changes in ocean circulation. However, the impact of WTIO diminishes toward the northern LMEs, such as the Mediterranean and Canary Currents, again reflecting the regional specificity of this index. A similar spatial pattern is observed for the Southwest Indian Ocean (SWIO) index in Fig. 10f, where sporadic positive regression coefficients are concentrated in the southwestern Indian Ocean. The Agulhas Current LME is particularly affected by SWIO, as this region is directly influenced by warming and mass redistribution in the southwestern Indian Ocean, mostly along the Madagascar coastal domain. In contrast, the Atlantic LMEs remain largely unaffected by SWIO, further emphasizing the spatial selectivity of Indian Ocean climate drivers. The influence of the central Pacific is captured by the Niño3.4 index in Fig. 10g, which shows a widespread impact across both the Indian Ocean and less impact on the Atlantic Ocean. Positive regression coefficients are evident in the western Indian Ocean close to the DMI patterns. This pattern reflects the teleconnections between the Pacific and other ocean basins, where changes in atmospheric circulation and oceanic processes driven by Pacific SST variability propagate their effects. The Somali and Red Sea Current LMEs are particularly sensitive to Niño3.4, while the Mediterranean and Agulhas LMEs remain less affected, indicating the reach but also the limits of Pacific influence. The Pacific Decadal Oscillation (PDO) (Fig. 10h) exhibits a notable large-scale influence on sea level variability, primarily across the Indian Ocean sector of the African LMEs. Positive regression coefficients are particularly evident in the Somali, Red Sea, and Agulhas Current LMEs, indicating that positive phases of the PDO typically associated with cooler SSTs in the western Pacific and warmer conditions in the eastern Pacific coincide with SLR in these regions. This response reflects enhanced Indian Ocean warming during positive PDO phases, which promotes upper-ocean thermal expansion and SLAs71,72. The Agulhas Current LME’s sensitivity may be linked to PDO-induced modulation of the Indian Ocean Walker circulation and subtropical gyre dynamics, influencing the transport and convergence of water masses along southeast Africa. In contrast, Atlantic LMEs, including the Guinea Current and Canary Current systems, show limited correlation with PDO, likely due to weaker atmospheric teleconnections between the Pacific and Atlantic basins in decadal timescales73. The Southern Annular Mode (SAM) (Fig. 10i), which reflects the north–south displacement of the westerly wind belt encircling Antarctica, shows no statistically significant impact on sea level trends across any African LMEs. This limited influence is consistent with previous findings that SAM-driven sea level variability is largely confined to southern extratropical latitudes, particularly the Southern Ocean and the high-latitude South Atlantic and Indian sectors74. As a result, the absence of significant regression patterns in African LMEs suggests that SAM’s dynamical effects on wind stress, Ekman transport, and associated mass redistribution do not strongly project onto the sea level variability of equatorial or subtropical African coastal systems.Fig. 10: Relationship between African sea level and major climate indices.Linear regression coefficients (mm) of sea level over Africa and its LMEs with respect to a ATL3, b TNA, c TASI, d DMI, e WTIO, f SWIO, g NINO 3.4, h PDO, and i SAM. White areas (white dots) indicate regions where regression coefficients are not statistically significant at the 95% confidence level.Full size imageSummary and discussionAccelerating global warming has intensified SLR along Africa’s coastal regions, creating unprecedented challenges for climate adaptation. Recent findings by Forster et al.75 reveal that the decade from 2014 to 2023 experienced an average global warming of 1.19 °C above pre-industrial levels, primarily due to anthropogenic emissions. In 2023 alone, the temperature anomaly reached 1.31 °C, further amplified by natural variability76. This unprecedented warming rate of 0.26 °C per decade stems from persistently high greenhouse gas emissions (~53 ± 5.4 Gt CO₂e annually) and a diminishing aerosol cooling effect75. These global climate trends are directly reflected in the SLAs across Africa’s seven LMEs, with 2023 marking a critical turning point for the continent.Between 1993 and 2023, mean sea levels across African LMEs rose by approximately 10.25 cm, equivalent to a linear trend of 3.31 ± 0.04 mm/yr. However, this average masks substantial spatial variability driven by complex interactions between thermosteric (temperature-induced), halosteric (salinity-induced), and manometric (mass-related) contributions. As shown in Table 1, all seven LMEs exhibited positive full-period trends (~2.70–3.91 mm/yr), yet the dominant physical drivers and their decadal evolution differ sharply across regions. In some basins (REDLME, BCLME), thermosteric expansion is the main contributor, while in others (MEDLME, GCLME, SCCLME), halosteric changes offset or modulate the warming signal, and since the early 2000s, manometric loading has surged to 2–3.6 mm/yr in most regions.The pace of change has accelerated dramatically over the past two decades, with global SLR rates reaching 3.80 ± 0.8 mm/yr from 2002 to 2023, exceeding the 3.52 ± 0.47 mm/yr observed over the full altimetric record (1993–2023). This recent acceleration is increasingly mass-driven in several key LMEs, such as the GCLME, where ocean mass addition has remained strong (~2.9–3.6 mm/yr since 2003), and the REDLME, where the last decade’s rise has been almost entirely steric and dominated by exceptional thermosteric gains (+3.23 mm/yr), with minimal halosteric offset due to stable salinity. These rising baselines elevate flood risks and erosion potential in low-lying deltas and urban zones, increasing the frequency of nuisance flooding and threatening coastal stability. The resulting physical impacts directly undermine SDG 14 (Life Below Water) by intensifying stress on marine habitats such as coral reefs, mangroves, and salt marshes, and SDG 11 (Sustainable Cities and Communities) by reducing the resilience of coastal settlements, infrastructure, and livelihoods against climate-driven sea-level hazards. At the heart of these trends lies Earth’s persistent energy imbalance (~0.9 W m⁻²), with oceans absorbing over 90% of excess heat driving both thermal expansion and accelerated land ice melt75,77. Recent research reveals that global ocean heat content has nearly doubled since 2010, with most heat accumulating in the upper 2000 m78. This pattern reflects the intensification of anthropogenic warming and its profound impact on ocean dynamics79. Consequently, the effects of these global processes manifest unevenly along Africa’s margins, creating distinct regional patterns of sea-level change.Nowhere is the temporal variability more striking than in several of Africa’s LMEs, particularly the REDLME in the east and the GCLME in the west, the two highest‑SLR LMEs on the continent, at 3.91 ± 0.13 mm/yr and 3.41 ± 0.05 mm/yr, respectively. Looking beyond these full-period averages, the REDLME shows one of the most abrupt decadal reversals anywhere along Africa’s coasts—shifting from a sharp SLA fall in 1993–2002 ( ~ − 7.14 mm/yr) to a rapid rise in 2003–2012 ( ~ + 6.18 mm/yr), and then to a still-positive, steric-dominated phase in 2013–2023 (+3.55 mm/yr, almost all thermosteric at +3.23 mm/yr). Similar stark regime shifts occurred in the ACLME (from −1.72 mm/yr to +1.01 mm/yr) and SCCLME (from −1.90 mm/yr to +5.92 mm/yr), highlighting coherent East African LME responses likely modulated by regional climate-ocean drivers such as the IOD and ENSO. In the REDLME, a strong and sustained thermosteric rise over the past two decades, largely unmitigated by salinity changes, aligns with its very high global coral-bleaching importance, since ongoing thermal ocean warming has increased the frequency and severity of marine heatwaves80—the proximate causes of coral bleaching in this region. In the GCLME, a mass-dominated rise has combined with halosteric contraction (salinification) to overwhelm vertical accretion in mangrove and salt marsh systems, raise hydroperiods in subsiding deltas, and exacerbate salinity stress, particularly in the Niger Delta, Lagos coast, and the barrier-lagoon complexes from Benin to Côte d’Ivoire. Negative halosteric trends here (−1.45 mm/yr in 2013–2023) imply increasing density and reduced stratification at depth, which can facilitate saline intrusion upriver during dry periods, stressing mangroves and estuarine marshes. Both regions are largely driven by upper-ocean warming7 process. In contrast, the MEDLME shows a more moderate increase of 2.70 ± 0.06 mm/yr, primarily linked to halosteric changes from high evaporation and limited freshwater input69, with recent strong salinification (−2.86 mm/yr halosteric) amplifying density gradients that inhibit vertical mixing, potentially reducing nutrient delivery to coastal marsh lagoons. Notably, mass redistribution from ice sheet melt now accounts for over 80% of total sea-level anomaly across Africa, with Greenland and Antarctica alone contributing about 2.0 mm/yr to the global mean75. According to Nicholls et al.81 high subsidence rates are widespread and not confined to a few well‑known locations. Multiple African coastal cities and deltaic regions, including Alexandria and Port Said on the Nile Delta; Abidjan, Cotonou, Lagos, and Douala along the Gulf of Guinea; and Mombasa and Maputo in East Africa, experience local land subsidence from groundwater extraction, sediment compaction, and infrastructure loading that adds several millimeters per year to climate‑driven SLR. In many of these locations, relative SLR now exceeds6–18 mm/yr58, substantially raising flood frequency and heightening exposure for millions of residents.The recent acceleration in SLR has been particularly pronounced across African oceanic regions, surging from 0.96 ± 0.26 mm/yr (1993–2002) to 4.34 ± 0.18 mm/yr (2013–2023), outpacing the global acceleration rate. For the ACLME, the negative SLA phase of 1993–2002 (−1.72 mm/yr) was driven primarily by a large negative manometric term (−3.10 mm/yr) despite a modestly positive thermosteric contribution, before rebounding to positive, mass‑dominated rises after 2003. SCCLME followed a similar pattern, emphasizing the east‑African coherency of these regime shifts. In several LMEs, including the GCLME, the shift from mixed steric–mass contributions in the 1990s to a clear mass-dominated signal after 2003 marks a structural change in the hazard baseline—one that is now decoupled from interdecadal thermosteric variability. This trend reflects a persistent global energy imbalance of approximately 0.9 W/m², with thermosteric expansion now contributing about 1.4 mm/yr to SLR75,82. The year 2023 proved particularly significant, with sea levels rising by 1.97 cm from 2022 to 2023, marking the second highest annual increase on record. The ACLME was especially affected, as intensified mesoscale activity and El Niño-related SST anomalies (+0.8 °C) generated regional sea-level hotspots7,83. These anomalies far exceeded the global average annual rise of 0.59 cm, posing acute risks to Africa’s low-lying coastal zones. Understanding the mechanisms driving these changes reveals the complexity of regional sea-level variability. These below-average trends align with regions of pronounced mesoscale eddy activity and complex current dynamics, particularly in the Mozambique Channel and southern Agulhas and Benguela systems84,85, where frequent eddy formation and southward migration influence sea-level variability through lateral heat redistribution and kinetic energy dissipation. Steric contributions varied significantly by region, with thermosteric and halosteric effects together accounting for 40–50% of observed sea-level changes, aligning with global patterns86. In the REDLME, thermosteric expansion in 2013–2023 alone reached +3.23 mm/yr, a physical signal closely tied to elevated bleaching risk in one of the planet’s most heat-tolerant coral provinces. In the GCLME, halosteric contraction has been consistently negative since 2003, possibly intensifying salinity intrusion in estuaries and compounding flood hazards from manometric loading. In 2023, thermosteric expansion in the REDLME reached 2.5 mm/yr, while marine heatwaves affected over 40% of the sea surface79,87. Halosteric effects were more pronounced in the GCLME and SCCLME, where enhanced rainfall from an intensified WAM caused SLA spikes of up to 8 cm, particularly near major river mouths like the Congo14,34. However, limited salinity data, especially in non-EBUS regions, constrains accurate quantification of halosteric trends, representing a critical observational gap. Manometric processes dominate in deltaic and subsiding areas such as the GCLME and CCLME, where ice melt and groundwater depletion have driven relative SLR to 18 mm/yr in places like Douala18 and Lagos20. Furthermore, large-scale climate oscillations significantly modulate these regional dynamics, adding another layer of complexity to sea-level variability.The 2023 El Niño, marked by strong Niño 3.4 SST anomalies, elevated sea levels in the SCCLME and triggered extreme rainfall that led to SLAs of up to 15 cm in the GCLME14. The Atlantic Niño intensified thermosteric rise in the GCLME and BCLME, with regression coefficients of 23.99 mm and 17.74 mm, respectively. Meanwhile, the Indian Ocean Dipole (DMI) played a key role in SLA variability in the ACLME and SCCLME, with positive phases increasing Congo River discharge by 20% and resulting in halosteric anomalies of up to 3 cm88. These climate-driven variations interact with the underlying trend of accelerating SLR to create compound hazards along African coastlines.The intersection of SLR and extreme weather events creates increasingly dangerous compound hazards that threaten coastal communities across Africa. Rising temperatures amplify the frequency and intensity of heatwaves, droughts, and heavy rainfall, threatening public health, food security, infrastructure, and labor productivity77. In 2023, El Niño-driven rains caused severe flooding in Ghana and southern Nigeria, regions already experiencing SLR of nearly 4 mm/yr89. Simultaneously, devastating floods in the Horn of Africa, following the worst drought in four decades, displaced hundreds of thousands and destroyed critical infrastructure90. Cyclone Freddy and Storm Daniel, which affected over 4 million people across Mozambique, Malawi, and Libya, were amplified by a 7 cm sea-level anomaly linked to Atlantic Meridional Overturning Circulation (AMOC) variability7,36. Such compound events increasingly exceed adaptation capacity, particularly among vulnerable populations where maternal and reproductive health face heightened risks55,77.These physical changes translate into profound socioeconomic consequences across the continent. More than 250 million people living in African coastal zones face escalating risks of chronic flooding, infrastructure collapse, and saltwater intrusion, with vulnerability particularly acute in megacities like Lagos, Alexandria, and Dar es Salaam91. In Lagos alone, subsidence may double flood frequency by 2050, threatening over 12 million residents15. Projections indicate that up to 117 million Africans could be affected by a 0.3-m SLR by 203092. This heightened vulnerability is compounded by rapid urbanization and inadequate coastal planning in many African cities, creating a cascade of interconnected risks.Ecological impacts are equally alarming across Africa’s marine ecosystems. In EBUS, like the Canary and Benguela Currents, warming reduces nutrient upwelling, undermining fishery productivity vital to food security25,93. In the REDLME and SCCLME, thermosteric warming accelerates mangrove dieback and coral bleaching, weakening natural coastal defences that protect communities from storm surges and erosion94. For island nations like Seychelles and Cape Verde, rising seas and intensifying storm surges threaten not only livelihoods and marine economies but also cultural heritage and national identity28,69. These ecological changes create feedback loops that further amplify human vulnerability, as degraded ecosystems provide less protection and fewer resources for coastal communities.The climate justice implications of these findings cannot be overstated. Africa’s negligible 4% contribution to global CO₂ emissions stands in stark contrast to the continent’s disproportionate suffering from climate consequences77. This disparity highlights the urgent need for international climate finance and technology transfer to support African adaptation efforts. Moreover, the accelerating pace of change, with SLR rates increasing from 0.96 mm/yr in the 1990s to over 4.3 mm/yr in recent years, suggests that current adaptation strategies may be insufficient for the challenges ahead.This comprehensive analysis reveals that SLR along Africa’s coastlines represents one of the most pressing climate challenges of our time. The strong spatial signature of manometric changes underscores the vulnerability of African coastal LMEs to global mass fluxes, which heighten risks of coastal flooding and erosion. The convergence of accelerating physical changes, intensifying extreme events, and growing coastal populations creates a perfect storm of vulnerability that demands immediate and sustained action. The scientific evidence presented here underscores the critical need for integrated coastal management strategies that address both the physical drivers of sea-level change and the socioeconomic vulnerabilities they exacerbate. As Africa continues to experience the frontline impacts of global climate change, the international community must recognize that supporting African adaptation efforts is not only a moral imperative but also essential for global climate stability and security. The time for incremental responses has passed; the scale and urgency of the challenge require transformative action that matches the magnitude of the threat facing Africa’s coastal regions.Data and methodologyThis study investigates the interannual variability and long-term trends in sea level within the Africa LMEs, focusing on steric (thermosteric and halosteric), manometric, and atmospheric contributions. A combination of satellite, in situ, and reanalysis products was employed, as outlined below.Satellite altimetry and climate indicesSLA were obtained from the Copernicus Marine and Environment Monitoring Service (CMEMS; DOI: 10.48670/moi-00148, accessed April 2025). The dataset, gridded at 0.25° × 0.25°, spans the period 1993–2023 and integrates observations from multiple satellite altimetry missions (Jason series, Sentinel-3A, HY-2A, Saral/AltiKa, CryoSat-2, TOPEX/Poseidon, ENVISAT, GFO, and ERS1/2). Standard corrections for instrumental biases, tidal effects, and dynamic atmospheric contributions (DAC) were applied. Specifically, DAC version 4.0, which includes wind effects and the inverted barometer (IB) correction, was used to isolate atmospheric influences on sea level variability95. Note that the SLA dataset had already been atmospherically corrected using this DAC product (AVISO; https://tds.aviso.altimetry.fr, accessed April 2025) and also was adjusted to account for drift in the TOPEX-A altimeter96 and the Jason-3 radiometer, which influences the wet troposphere correction. To correct for GIA the viscoelastic response of the Earth to deglaciation and associated ocean basin redistribution, geoid height changes rates from the ICE6G-D model97 were resampled to match the spatial resolution of the altimetry data.Monthly time series of relevant climate indices (Supplementary Fig. 4) were used to examine the influence of large-scale climate variability on regional sea level.These include:Niño 3.4 IndexRepresents central equatorial Pacific variability and is calculated as the area-averaged sea surface temperature (SST) anomaly over 170°W–120°W and 5°S–5°N. Positive anomalies are associated with El Niño events, typically peaking during boreal fall and winter.Tropical North Atlantic (TNA) IndexCaptures temperature anomalies in the North Atlantic region (55°W–15°W, 5°N–25°N), which influence atmospheric circulation and precipitation over the adjacent continents.Tropical South Atlantic (TSA) IndexAnalogous to the TNA, this index reflects anomalies in the South Atlantic and is used in conjunction with TNA to identify a dipolar mode of variability in the Atlantic sector98.Tropical Atlantic SST Gradient Index (TASI)Defined as the difference between the TNA and TSA indices, TASI captures the meridional gradient of sea surface conditions in the Atlantic, a key indicator of the Atlantic dipole mode98.Southwest Indian Ocean (SWIO) IndexRepresents variability in the region southeast of Madagascar (31°E–45°E, 32°S–25°S), where elevated temperatures have been linked to enhanced austral summer rainfall over southern Africa99.Western Tropical Indian Ocean (WTIO) IndexReflects SST variability in the western Indian Ocean (50°E–70°E, 10°S–10°N) and contributes to the Dipole Mode Index (DMI) calculated as the difference between WTIO and the southeastern Indian Ocean (SETIO: 90°E–110°E, 10°S–0°S) a key metric of the Indian Ocean Dipole100. High DMI values have been associated with anomalous East African rainfall101.Atlantic 3 (ATL3) IndexRepresents SST anomalies in the equatorial Atlantic (20°W–0°, 2.5°S–2.5°N) and is used to monitor equatorial Atlantic variability based on a ± 0.4 °C threshold.Pacific Decadal Oscillation (PDO)A low-frequency climate pattern with El Niño-like manifestations primarily in the North Pacific and North American regions.Southern Annular Mode (SAM)Also known as the Antarctic Oscillation, the SAM index describes variations in the zonal pressure gradient between the mid-latitudes and Antarctica, influencing the climate of the Southern Hemisphere.All climate indices were sourced from the NOAA Physical Sciences Laboratory (https://psl.noaa.gov/gcos_wgsp/Timeseries, accessed April 2025).Steric sea levelSteric sea level variations and their components were derived from the CMEMS ARMOR-3D product (https://doi.org/10.48670/moi-00052, accessed April 2025). This dataset provides monthly temperature and salinity profiles at a spatial resolution of 0.25° × 0.25° and 50 vertical levels, covering the period 1993–2023. The fields are generated through optimal interpolation of in situ measurements (e.g., CTDs, ARGO floats, XBTs) and satellite observations and are widely recognized for robust steric sea level analyses102.The steric component of sea level anomalies (SSLA) is calculated using the formula14,34,103,104:$${SSLA} = – int^{0}_{-H}frac{Delta rho }{{rho }_{0}}{dz}$$where ({rho }_{0}) = 1025 kg/m3 is the reference density, (Delta rho) is the density anomaly with relative to the 1993–2023 climatology, and H represents the integration depth (either 700 m or the local seafloor, whichever is shallower).Thermosteric (TSLA) and halosteric (HSLA) components were estimated separately using the TEOS-10 formulation105,106,107:$${TSLA} = int_{-H}^{0}alpha Delta T{dz},, {HSLA} = – int_{-H}^{0}beta Delta S{dz}$$Here, (alpha) and (beta) denote the thermal expansion and haline contraction coefficients, respectively, which were computed in accordance with the Thermodynamic Equation of Seawater standard105. The variables (Delta T) and (Delta S) are the anomalies in temperature and salinity relative to the same climatological baseline. All calculations were limited to the upper 700 m, as this layer is well-observed and captures the dominant steric signal over the LMEs’ typically shallow continental shelves. Contributions below 700 m were excluded due to their minor role in interannual variability and higher uncertainty108. Ocean mass (Manometric sea level) changes were inferred as the residual between total SLA (from Section “Regional sea level rise and the exceptional 2023 event in Africa”) and steric SLA (from this section), with the SLA field already corrected for GIA39, and no further correction applied to the residual. To isolate non-seasonal signals, seasonal cycles (annual and semi-annual) were removed at each grid point using a least-squares fit of 12- and 6-month sinusoids. A 3-month Lanczos filter was then applied to each dataset to suppress high-frequency variability. Long-term linear trends were derived from the deseasonalized, filtered monthly time series using ordinary least squares regression109. Uncertainties were expressed as 95% confidence intervals (±2σ) and corrected for serial correlation110 to ensure robust error estimates. To assess statistical significance, we used the Modified Mann–Kendall test111, a non-parametric method that accounts for both autocorrelation and non-normality in the time series, providing a rigorous evaluation of trend reliability.

    Data availability

    Copernicus Marine Environment Monitoring Service (CMEMS) altimetry products are publicly available at https://doi.org/10.48670/moi-00148. The CMEMS ARMOR-3D temperature and salinity fields are available at https://doi.org/10.48670/moi-00052. AVISO Dynamic Atmospheric Correction (DAC v4.0) products can be accessed at https://tds.aviso.altimetry.fr. Climate indices (Niño 3.4, TNA, TSA, TASI, SWIO, WTIO, SETIO, DMI, ATL3, PDO, SAM) are publicly available from the NOAA Physical Sciences Laboratory at https://psl.noaa.gov/gcos_wgsp/Timeseries. The ICE6G-D model outputs used for glacial isostatic adjustment (GIA) corrections are available from Peltier et al.97.
    Code availability

    The codes112 utilized to produce the findings in this article can be found at https://doi.org/10.5281/zenodo.17401591. The color schemes adhere to the color-blind-friendly palettes of Crameri et al.113, and the mapping was done using the Generic Mapping Tools (GMT)114.
    ReferencesLeatherman, S. P. Social and economic costs of sea level rise. in International Geophysics (Elsevier, 2001).Rullens, V. et al. Understanding the consequences of sea level rise: the ecological implications of losing intertidal habitat. N. Zeal J. Mar. Freshw. Res. 56, 353–370 (2022).Article 

    Google Scholar 
    WMO. State of the Global Climate 2023. (2024). https://doi.org/10.18356/9789263113474 (2025).WMO. State of the Climate in Africa 2024. 7 https://doi.org/10.18356/9789263113702 (2025).Cazenave, A. & Moreira, L. Contemporary sea-level changes from global to local scales: a review. Proc. R. Soc. A 478, 20220049 (2022).Article 

    Google Scholar 
    Durand, G. et al. Sea-level rise: from global perspectives to local services. Front. Mar. Sci. 8 https://doi.org/10.3389/fmars.2021.709595 (2022).Minobe, S. et al. Global and regional drivers for exceptional climate extremes in 2023-2024: beyond the new normal. NPJ Clim. Atmos. Sci. 8, 138–111 (2025).Article 

    Google Scholar 
    Ripple, W. J. et al. The 2023 state of the climate report: entering uncharted territory. Bioscience 73, 841–850 (2023).Article 

    Google Scholar 
    Kemgang Ghomsi, F. E. et al. Sea level trends along the South African coast from 1993 to 2022 using XTRACK altimetry, tide gauges, and GNSS measurements. Sci. Rep. 15, 4946–4915 (2025).Article 
    CAS 

    Google Scholar 
    Kemgang Ghomsi, F. E. et al. Vertical crustal movement along the coast of South Africa. Int. Arch. Photogramm. Remote Sens. Spat. Inf. Sci. 48, 393–397 (2025).Article 

    Google Scholar 
    Trisos, C. et al. Africa. In: Climate Change 2022: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change (Cambridge University Press, 2023).Maul, G. A., Duedall, I. W., Makowski, C. & Finkl, C. W. Encyclopedia of Earth Sciences Series (Springer International Publishing, 2019).Aretouyap, Z., Ghomsi, F. E. K., Domra, J. K., Tchato, S. C. & Perilli, N. Assessing groundwater potential in Foumban and implication on urban planning in the context of sustainable development. Geocarto Int. 38 https://doi.org/10.1080/10106049.2022.2163307 (2023).Ghomsi, F. E. K. et al. Exploring steric sea level variability in the Eastern Tropical Atlantic Ocean: a three-decade study (1993–2022). Sci. Rep. 14, 20458 (2024).Article 
    CAS 

    Google Scholar 
    Ghomsi, F. E. K. et al. Sea level rise and coastal flooding risks in the Gulf of Guinea. Sci. Rep. 14, 1–11 (2024).Article 

    Google Scholar 
    Nhantumbo, B. J., Dada, O. A. & Ghomsi, F. E. Sea level rise and climate change-impacts on African coastal systems and cities. https://doi.org/10.5772/intechopen.113083 (2023).Njiiri, W., Njuguna, M. & Wahome, E. Establishing historical shoreline trends for adaption of maritime built heritage to climate change part 2: the Northern Kenya Coast. Hist. Environ. 16, 167–186 (2025).Article 

    Google Scholar 
    Chounna Yemele, G. et al. Land subsidence dynamics and their interplay with spatial and temporal land-use transitions in the Douala coastland, Cameroon. EGUsphere 2025, 1–29 (2025).
    Google Scholar 
    Daramola, O. E. et al. Detection of crustal deformation using a gravimetric and Multitemporal Interferometric SAR (MT-InSAR) approach: detection of crustal deformation. Pure Appl. Geophys. 182, 2111–2134 (2025).Article 

    Google Scholar 
    Ohenhen, L. & Shirzaei, M. Land subsidence hazard and building collapse risk in the coastal city of Lagos, West Africa. Earth’s Future 10, e2022EF003219 (2022).Article 
    CAS 

    Google Scholar 
    Magalhães Filho, L. N. L., Roebeling, P. C., Costa, L. F. C. & de Lima, L. T. Ecosystem services values at risk in the Atlantic coastal zone due to sea-level rise and socioeconomic development. Ecosyst. Serv. 58, 101492 (2022).Article 

    Google Scholar 
    Sweijd, N. A. & Smit, A. J. Trends in sea surface temperature and chlorophyll-a in the seven African large marine ecosystems. Environ. Dev. 36, 100585 (2020).Article 

    Google Scholar 
    Ibrahim, A., Salifu, A.-H. & Peprah, C. Does governance matter when disaster looms? Zooming into proactive institutional measures for flood risk management. Int. J. Disaster Risk Reduct. 97, 104021 (2023).Article 

    Google Scholar 
    Sarkodie, S. A., Ahmed, M. Y. & Owusu, P. A. Global adaptation readiness and income mitigate sectoral climate change vulnerabilities. Human. Soc. Sci. Commun. 9, 1–17 (2022).
    Google Scholar 
    Bakun, A. et al. Anticipated effects of climate change on coastal upwelling ecosystems. Curr. Clim. Change Rep. 1, 85–93 (2015).Article 

    Google Scholar 
    Ramírez, F., Shannon, L. J., Angelini, R., Steenbeek, J. & Coll, M. Overfishing species on the move may burden seafood provision in the low-latitude Atlantic Ocean. Sci. total Environ. 836, 155480–155480 (2022).Article 

    Google Scholar 
    Jury, C. P. & Toonen, R. J. Widespread scope for coral adaptation under combined ocean warming and acidification. Proc. R. Soc. B 291, 20241161 (2024).Article 

    Google Scholar 
    Nurse, L. A. et al. Small islands. Climate change 2014: Impacts, adaptation, and vulnerability. Part B: Regional aspects. Contribution of working group II to the fifth assessment report of the intergovernmental panel on climate change, pp. 1613-1654 (2014).Tamoffo, A. T., Weber, T. & Saleem, F. Population exposed to compound extreme hot & humid and wet & windy events in Africa. Environ. Res. Lett. 20, 14038 (2025).Article 

    Google Scholar 
    Didi, S. R., Diakhaté, M. & Diedhiou, A. Changes in the West Africa Monsoon precipitation extremes during ENSO developing Phases. Atmos.-ocean 61, 293–305 (2023).Article 

    Google Scholar 
    Aryee, J. N. A. et al. Quantifying climatic heavy-precipitation-induced floods in West Africa using multiple precipitation indices Quantifying climatic heavy-precipitation-induced floods in West Africa using multiple precipitation indices. Sci. Afr. 25, e02309 (2024).
    Google Scholar 
    Kenfack, K. et al. Radiative anomalies associated with extreme precipitation of November 2023 in equatorial Central Africa. Atmos. Res. 321, 108090 (2025).Article 

    Google Scholar 
    Tchana, B. C. et al. Analysis of air temperature anomalies during abnormal rainfall events in Western Central Africa from September to November 2023: Analysis of air temperature anomalies during abnormal rainfall events in Western Central Africa from September to November 2023. Theor. Appl. Climatol. 156, 288 (2025).Article 

    Google Scholar 
    Ghomsi, F. E. K. et al. Sea level variability in Gulf of Guinea from satellite altimetry. Sci. Rep. 14, 4759 (2024).Article 

    Google Scholar 
    Saber, M. et al. A comprehensive review of climate change adaptation and disaster risk reduction in Africa. J. Water Clim. change 16, 1831–1862 (2025).Article 

    Google Scholar 
    Carbon, B. Analysis: Africa’s extreme weather has killed at least 15,000 people in 2023. Carbon Brief. https://www.carbonbrief.org/analysis-africas-extreme-weather-have-killed-at-least-15000-people-in-2023/ (2023).Zaine, P., Ramontsheng, R., Sarah, R. & Moagabo, R. Analysing the atmospheric-oceanic conditions driving the sustained long track and intensity of Tropical Cyclone Freddy Analysing the atmospheric-oceanic conditions driving the sustained long track and intensity of Tropical Cyclone Freddy. Trop. Cyclone Res. Rev. 13, 356–388 (2024).Article 

    Google Scholar 
    Armon, M., Shmilovitz, Y. & Dente, E. Anatomy of a foreseeable disaster: Lessons from the 2023 dam-breaching flood in Derna, Libya. Sci. Adv. 11, eadu2865 (2025).Article 

    Google Scholar 
    Cazenave, A. et al. Global sea-level budget 1993–present. Earth Syst. Sci. Data 10, 1551–1590 (2018).Article 

    Google Scholar 
    Lian, T., Wang, J., Chen, D., Liu, T. & Wang, D. A strong 2023/24 El Niño is Staged by Tropical Pacific Ocean Heat Content Buildup. Ocean-land-atmos. Res. 2, 187–193 (2023).Article 

    Google Scholar 
    Peng, Q., Xie, S.-P., Passalacqua, G. A., Miyamoto, A. & Deser, C. The 2023 extreme coastal El Niño: atmospheric and air-sea coupling mechanisms. Sci. Adv. 10, eadk8646 (2024).Article 

    Google Scholar 
    Ericson, J. P., Vörösmarty, C. J., Dingman, S. L., Ward, L. G. & Meybeck, M. Effective sea-level rise and deltas: causes of change and human dimension implications. Glob. Planet. Change 50, 63–82 (2006).Article 

    Google Scholar 
    Nicholls, R. J. Coastal flooding and wetland loss in the 21st century: changes under the SRES climate and socio-economic scenarios. Glob. Environ. Change 14, 69–86 (2004).Article 

    Google Scholar 
    Leclercq, L. et al. Spatio-temporal changes in interannual sea level along the world coastlines. Glob. Planet. Change, 104972 https://doi.org/10.1016/j.gloplacha.2025.104972 (2025).Intergovernmental Panel on Climate, C. Climate Change 2022 – Impacts, Adaptation and Vulnerability: Working Group II Contribution to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change (Cambridge University Press, 2023).Rounce, D. R. et al. Downstream hydrology reduces glaciers’ direct contribution to sea-level rise. Geophys. Res. Lett. 52, e2025GL114866 https://doi.org/10.1029/2025GL114866 (2025).Parkes, D. & Marzeion, B. Twentieth-century contribution to sea-level rise from uncharted glaciers. Nature 563, 551–554 (2018).Article 
    CAS 

    Google Scholar 
    Shi, Q., Pavey, E. S. & Carter, R. E. Bonferroni-based correction factor for multiple, correlated endpoints. Pharm. Stat. 11, 300–309 (2012).Article 

    Google Scholar 
    Hamlington, B. et al. The rate of global sea level rise doubled during the past three decades. Commun. Earth Environ. 5, 601 (2024).Article 

    Google Scholar 
    Bellas-Manley, A., Nerem, R. S. & Hamlington, B. D. Extrapolation of the Satellite Altimeter record to understand regional variations in future sea level change. J. Geophys. Res. Oceans 130, e2024JC022094 https://doi.org/10.1029/2024JC022094 (2025).Mitrovica, J. X. et al. On the robustness of predictions of sea level fingerprints. Geophys. J. Int. 187, 729–742 (2011).Article 

    Google Scholar 
    Dussaillant, I. et al. Annual mass change of the world’s glaciers from 1976 to 2024 by temporal downscaling of satellite data with in situ observations. Earth Syst. Sci. Data https://doi.org/10.3929/ethz-b-000735850 (2025).von Schuckmann, K. et al. Heat stored in the Earth system 1960–2020: where does the energy go? Earth Syst. Sci. Data 15, 1675–1709 (2023).Article 

    Google Scholar 
    Pörtner, H.-O. et al. The ocean and cryosphere in a changing climate. IPCC special report on the ocean and cryosphere in a changing climate. Cambridge University Press 1155 https://doi.org/10.1017/9781009157964 (2019).Intergovernmental Panel on Climate Change (IPCC) (Ed.). Climate Change 2022—Mitigation of Climate Change https://doi.org/10.1017/9781009157926 (2023).Spalding, M. D. et al. Coastal ecosystems: a critical element of risk reduction. Conserv. Lett. 7, 293–301 (2014).Article 

    Google Scholar 
    Frederikse, T. et al. The causes of sea-level rise since 1900. Nature 584, 393–397 (2020).Article 
    CAS 

    Google Scholar 
    Syvitski, J. P. M. et al. Sinking deltas due to human activities. Nat. Geosci. 2, 681–686 (2009).Article 
    CAS 

    Google Scholar 
    Hutchings, L. et al. The Benguela current: an ecosystem of four components. Prog. Oceanogr. 83, 15–32 (2009).Article 

    Google Scholar 
    Durgadoo, J. V., Rühs, S., Biastoch, A. & Böning, C. W. B. Indian Ocean sources of Agulhas leakage. J. Geophys. Res. Oceans 122, 3481–3499 (2017).Article 

    Google Scholar 
    Llovel, W. et al. Global ocean freshening, ocean mass increase and global mean sea level rise over 2005–2015. Sci. Rep. 9, 17717 (2019).Article 

    Google Scholar 
    Zemp, M. et al. Community estimate of global glacier mass changes from 2000 to 2023. Nature 639, 382–388 (2025).Article 

    Google Scholar 
    Swapna, P., Jyoti, J., Krishnan, R., Sandeep, N. & Griffies, S. M. Multidecadal weakening of indian summer monsoon circulation induces an increasing Northern Indian Ocean Sea Level. Geophys. Res. Lett. 44, 10,560–510,572 (2017).Article 

    Google Scholar 
    Chatterjee, A., Kumar, B. P., Prakash, S. & Singh, P. Annihilation of the Somali upwelling system during summer monsoon. Sci. Rep. 9, 7598–7598 (2019).Article 
    CAS 

    Google Scholar 
    Stüwe, K., Robl, J., Turab, S. A., Sternai, P. & Stuart, F. M. Feedbacks between sea-floor spreading, trade winds and precipitation in the Southern Red Sea. Nat. Commun. 13, 5405–5405 (2022).Article 

    Google Scholar 
    Bonaduce, A., Pinardi, N., Oddo, P., Spada, G. & Larnicol, G. Sea-level variability in the Mediterranean Sea from altimetry and tide gauges. Clim. Dyn. 47, 2851–2866 (2016).Article 

    Google Scholar 
    Legeais, J.-F. et al. An improved and homogeneous altimeter sea level record from the ESA Climate Change Initiative. Earth Syst. Sci. Data 10, 281–301 (2018).Article 

    Google Scholar 
    Marcos, M. & Tsimplis, M. N. Coastal sea level trends in Southern Europe. Geophys. J. Int. 175, 70–82 (2008).Article 

    Google Scholar 
    Borile, F. et al. The Eastern Mediterranean Sea mean sea level decadal slowdown: the effects of the water budget. Front. Clim. 7 https://doi.org/10.3389/fclim.2025.1472731 (2025).Durack, P. J., Wijffels, S. E. & Matear, R. J. Ocean salinities reveal strong global water cycle intensification during 1950 to 2000. Science 336, 455–458 (2012).Article 
    CAS 

    Google Scholar 
    Han, W. et al. Spatial patterns of sea level variability associated with natural internal climate modes. Surv. Geophys. 38, 217–250 (2017).Article 

    Google Scholar 
    Zhang, X. & Church, J. A. Sea level trends, interannual and decadal variability in the Pacific Ocean. Geophys. Res. Lett. 39, L21701 (2012).Article 

    Google Scholar 
    Meehl, G. A. et al. Atlantic and Pacific tropics connected by mutually interactive decadal-timescale processes. Nat. Geosci. 14, 36–42 (2021).Article 
    CAS 

    Google Scholar 
    Boening, C., Willis, J. K., Landerer, F. W., Nerem, R. S. & Fasullo, J. The 2011 La Niña: so strong, the oceans fell. Geophys. Res. Lett. 39, L19602 (2012).Article 

    Google Scholar 
    Forster, P. M. et al. Indicators of Global Climate Change 2023: annual update of key indicators of the state of the climate system and human influence. Earth Syst. Sci. Data 16, 2625–2658 (2024).Article 

    Google Scholar 
    Bevacqua, E., Schleussner, C.-F. & Zscheischler, J. A year above 1.5 °C signals that Earth is most probably within the 20-year period that will reach the Paris Agreement limit. Nat. Clim. change 15, 262–265 (2025).Article 

    Google Scholar 
    Schaeffer, R., Gustafsson, Ö., Riipinen, I. & Ziervogel, G. Ten new insights in climate science 2024. One Earth https://doi.org/10.1016/j.oneear.2025.101285 (2025).Li, Z., England, M. H. & Groeskamp, S. Recent acceleration in global ocean heat accumulation by mode and intermediate waters. Nat. Commun. 14, 6888–6814 (2023).Article 
    CAS 

    Google Scholar 
    Cheng, L. et al. Record high temperatures in the Ocean in 2024. Adv. Atmos. Sci. 42, 1092–1109 (2025).Article 

    Google Scholar 
    Huang, B. et al. Extreme marine heatwaves in the global oceans during the past decade. Bull. Am. Meteorol. Soc.https://doi.org/10.1175/BAMS-D-24-0337.1 (2025).Nicholls, R. J. et al. A global analysis of subsidence, relative sea-level change and coastal flood exposure. Nat. Clim. Change 11, 338–342 (2021).Article 

    Google Scholar 
    Hakuba, M. Z. et al. Trends and Variability in Earth’s energy imbalance and ocean heat uptake since 2005. Surv. Geophys. 45, 1721–1756 (2024).Article 

    Google Scholar 
    Jiang, N. et al. El Niño and sea surface temperature pattern effects lead to historically high global mean surface temperatures in 2023. Geophys. Res. Lett. 52, e2024GL113733 https://doi.org/10.1029/2024GL113733 (2025).Matano, R. P. & Palma, E. D. On the upwelling of downwelling currents. J. Phys. Oceanogr. 38, 2482–2500 (2008).Article 

    Google Scholar 
    Swart, N. C., Lutjeharms, J. R. E., Ridderinkhof, H. & de Ruijter, W. P. M. Observed characteristics of Mozambique Channel eddies. J. Geophys. Res. Atmos. 115 https://doi.org/10.1029/2009JC005875 (2010).Fox-Kemper, B. H. et al. 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change (eds Zhai Masson-Delmotte, P. et al.) 1211–1362 (Cambridge University Press, 2021).Gunnarson, J. L. et al. Removing ENSO’s influence from global SST variability, with insights into the record-setting marine heatwaves of 2023-2024. Bull. Am. Meteorol. Soc. https://doi.org/10.1175/BAMS-D-24-0023.1 (2025).Jarugula, S. & McPhaden, M. J. Indian Ocean Dipole affects eastern tropical Atlantic salinity through Congo River Basin hydrology. Commun. Earth Environ. 4, 366–369 (2023).Article 

    Google Scholar 
    West Africa Seasonal, M. 2023 Season – August Update. (https://reliefweb.int/attachments/5a81cb98-60b6-4c00-a60d-751bb3d62cd3/WFP-0000151903.pdf, 2023).Cai, W. et al. Climate impacts of the El Niño–Southern Oscillation in Africa. Nat. Rev. Earth Environ. 6, 503–520 (2025).Article 

    Google Scholar 
    United Nations Human Settlements, P. State of the world’s cities, 2008/2009: harmonious cities. (2008).Bindoff, N. L. et al. Changing ocean, marine ecosystems, and dependent communities. Intergovernmental Panel on Climate Change (IPCC), Intergovernmental Panel https://doi.org/10.1017/9781009157964.007 (2022).García-Reyes, M. et al. Under Pressure: Climate Change, Upwelling, and Eastern Boundary Upwelling Ecosystems. Front. Mar. Sci. 2 https://doi.org/10.3389/fmars.2015.00109 (2015).Dubois, K., Larsen, M. A. D., Drews, M., Nilsson, E. & Rutgersson, A. Influence of data source and copula statistics on estimates of compound flood extremes in a river mouth environment. Nat. Hazards Earth Syst. Sci. 24, 3245–3265 (2024).Article 

    Google Scholar 
    Landerer, F. W. & Volkov, D. L. The anatomy of recent large sea level fluctuations in the Mediterranean Sea. Geophys. Res. Lett. 40, 553–557 (2013).Article 

    Google Scholar 
    Ablain, M. et al. The TOPEX-A drift and impacts on GMSL time series Comparison with Poseïdon, p. 2017, https://meetings.aviso.altimetry.fr/fileadmin/user_ upload/tx_ausyclsseminar/files/Poster_OSTST17_GMSL_ Drift_TOPEX-A.pdf (2017). (last access: 2 November 2022).Peltier, W. R., Argus, D. F. & Drummond, R. Comment on “An assessment of the ICE-6G_C (VM5a) glacial isostatic adjustment model” by Purcell et al. J. Geophys. Res. Solid Earth 123, 2019–2028 (2018).Article 

    Google Scholar 
    Enfield, D. B., Mestas-Nuñez, A. M. & Trimble, P. J. The Atlantic multidecadal oscillation and its relation to rainfall and river flows in the continental US. Geophys. Res. Lett. 28, 2077–2080 (2001).Article 

    Google Scholar 
    Hermes, J. C. & Reason, C. J. C. Ocean model diagnosis of interannual coevolving SST variability in the South Indian and South Atlantic Oceans. J. Clim. 18, 2864–2882 (2005).Article 

    Google Scholar 
    Saji, N., Goswami, B. N., Vinayachandran, P. & Yamagata, T. A dipole mode in the tropical Indian Ocean. Nature 401, 360–363 (1999).Article 
    CAS 

    Google Scholar 
    Black, E., Laughton, A. S., Flemming, N. C. & Spencer, T. The relationship between Indian Ocean sea–surface temperature and East African rainfall. Philos. Trans. R. Soc. Lond. Ser. A: Math. Phys. Eng. Sci. 363, 43–47 (2005).
    Google Scholar 
    Gasparin, F. et al. Requirements for an integrated in situ Atlantic Ocean observing system from coordinated observing system simulation experiments. Front. Mar. Sci. 6 https://doi.org/10.3389/fmars.2019.00083 (2019).Siegismund, F., Johannessen, J., Drange, H., Mork, K. A. & Korablev, A. Steric height variability in the Nordic Seas. J. Geophys. Res. C. Oceans 112, n/a https://doi.org/10.1029/2007JC004221 (2007).Piecuch, C. G. & Ponte, R. M. Mechanisms of Global-mean steric sea level change. J. Clim. 27, 824–834 (2014).Article 

    Google Scholar 
    Pawlowicz, R. & Feistel, R. Limnological applications of the Thermodynamic Equation of Seawater 2010 (TEOS-10). Limnol. Oceanogr. Methods 10, 853–867 (2012).Article 
    CAS 

    Google Scholar 
    Jayne, S. R., Wahr, J. M. & Bryan, F. O. Observing ocean heat content using satellite gravity and altimetry. J. Geophys. Res. C. Oceans 108, 13.11-n/a https://doi.org/10.1029/2002JC001619 (2003).Wang, G., Cheng, L., Boyer, T. & Li, C. Halosteric sea level changes during the Argo era. Water 9, 484 (2017).Article 

    Google Scholar 
    Purkey, S. G. & Johnson, G. C. Warming of Global Abyssal and Deep Southern Ocean Waters between the 1990s and 2000s: contributions to global heat and sea level rise budgets. J. Clim. 23, 6336–6351 (2010).Article 

    Google Scholar 
    Wooditch, A., Johnson, N. J., Solymosi, R., Medina Ariza, J. & Langton, S. Ordinary Least Squares Regression (Springer International Publishing, 2021).Emery, W. J., Talley, L. D. & Pickard, G. L. Descriptive physical oceanography. available at: HTU http://www-pord.ucsd.edu/~ltalley/sio210/pickard_emery/UT (2006).Alashan, S. Combination of modified Mann-Kendall method and Şen innovative trend analysis. Engineering Rep. 2, n/a https://doi.org/10.1002/eng2.12131 (2020).Ghomsi, F. E. K., Stroeve, J., Bonaduce, A. & Raj, R. P. Accelerating sea level rise in Africa and its large marine ecosystems since the 1990s. Zenodo. https://doi.org/10.5281/zenodo.17401590 (2025).Crameri, F., Shephard, G. & Heron, P. J. Choosing suitable color palettes for accessible and accurate science figures. https://doi.org/10.1002/cpz1.1126 (2024).Wessel, P. et al. The generic mapping tools version 6. Geochem. Geophys. Geosyst. 20, 5556–5564 (2019).Article 

    Google Scholar 
    Download referencesAcknowledgementsF.E.K.G. and J.S. acknowledge funding from Canada’s C150 Research Program (Grant No. 50296). F.E.K.G., R.P.R., and A.B. are supported by the Sea Level Predictions and Reconstructions (SeaPR) project, funded through the Bjerknes Centre for Climate Research (BCCR) Strategic Projects Initiative. F.E.K.G received support from Schmidt Sciences, LLC. R.P.R. and A.B. also acknowledge support from the European Space Agency’s DRAGON 5 project. We are grateful to the three anonymous reviewers, the Editorial Board Member, and the Deputy Editor for their thoughtful and constructive feedback, which greatly improved this work.Author informationAuthors and AffiliationsCentre for Earth Observation Science, University of Manitoba, Winnipeg, MB, CanadaFranck Eitel Kemgang Ghomsi & Julienne StroeveDepartment of Oceanography, University of Cape Town, Cape Town, South AfricaFranck Eitel Kemgang GhomsiGeodesy Research Laboratory, National Institute of Cartography, Yaoundé, CameroonFranck Eitel Kemgang GhomsiDepartment of Earth Sciences, University College London, London, UKJulienne StroeveNational Snow and Ice Data Center, Cooperative Institute for Research in Environmental Science, University of Colorado Boulder, Boulder, CO, USAJulienne StroeveNansen Environmental and Remote Sensing Center and Bjerknes Center for Climate Research, Bergen, NorwayAntonio Bonaduce & Roshin P. RajAuthorsFranck Eitel Kemgang GhomsiView author publicationsSearch author on:PubMed Google ScholarJulienne StroeveView author publicationsSearch author on:PubMed Google ScholarAntonio BonaduceView author publicationsSearch author on:PubMed Google ScholarRoshin P. RajView author publicationsSearch author on:PubMed Google ScholarContributionsF.E.K.G. conceived the study, led the analysis, and developed the methodology, software, and visualizations. J.S. contributed to the methodology, validation, and provided overall supervision. A.B. and R.P.R. assisted with methodological development and validation. Project management was jointly carried out by F.E.K.G. and J.S. F.E.K.G. drafted the manuscript, and all authors contributed to its revision and approved the final version.Corresponding authorCorrespondence to
    Franck Eitel Kemgang Ghomsi.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Peer review

    Peer review information
    Communications Earth & Environment thanks the anonymous reviewers for their contribution to the peer review of this work. Primary Handling Editors: Olusegun Dada, Joe Aslin and Alice Drinkwater. A peer review file is available.

    Additional informationPublisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary informationTransparent Peer Review fileSupplementary InformationRights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleGhomsi, F.E.K., Stroeve, J., Bonaduce, A. et al. Accelerating sea level rise in Africa and its large marine ecosystems since the 1990s.
    Commun Earth Environ 6, 1008 (2025). https://doi.org/10.1038/s43247-025-02965-zDownload citationReceived: 14 July 2025Accepted: 29 October 2025Published: 15 December 2025Version of record: 15 December 2025DOI: https://doi.org/10.1038/s43247-025-02965-zShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Twentieth Century Reanalysis version 3 as a source of information on long-term trends (1806–2022) in lake surface water temperature changes in Central Europe (Poland)

    AbstractWater temperature is one of the fundamental characteristics of the hydrosphere, determining the functioning of its various components. In the case of lakes, surface water temperature shows a strong correlation with air temperature, and this relationship forms the basis for reconstructing the thermal regime of lakes. The study uses the Twentieth Century Reanalysis (20CRv3) meteorological dataset to reconstruct the surface water temperature of seven lakes in Poland for the period 1806–2022. This approach significantly expands the current state of knowledge, particularly for Central Europe, and includes periods predating significant human impact on the environment. Over the course of more than 200 years, an increase in water temperature has been observed, averaging 0.081 °C per decade across all studied lakes. Considering the changes in water temperature in the studied lakes, several distinct phases can be observed, which generally reflect changes in climatic conditions. Based on the results of the Pettitt test, the characteristic points include the 1840s, the 1940s, and the late 1980s. Rapid warming has been recorded in recent decades, and current studies suggest this trend is likely to continue in the future. This situation calls for multidisciplinary consultation and subsequent action to develop strategies for mitigating the impact of global warming on lake ecosystems.

    Similar content being viewed by others

    Global lakes are warming slower than surface air temperature due to accelerated evaporation

    Article

    23 October 2023

    The impact of extreme heat on lake warming in China

    Article
    Open access
    02 January 2024

    Seasonal overturn and stratification changes drive deep-water warming in one of Earth’s largest lakes

    Article
    Open access
    16 March 2021

    IntroductionKnowledge about the natural environment depends on the duration and accuracy of measurements related to its individual components. One of the fundamental characteristics of the atmosphere is air temperature, which defines its thermal state and determines the rate and scale of many processes and phenomena occurring within the Earth’s climate system. Therefore, with the progress of civilization, instrumental measurements of air temperature were undertaken relatively early compared to other environmental studies1,2,3. Linking air temperature—and other climatic elements characterizing the atmosphere (precipitation, wind)—with other components of the environment (geosphere, biosphere, hydrosphere) enables the assessment of changes in those components over centuries. Particularly strong relationships are observed between the atmosphere and the hydrosphere. These connections are used to reconstruct gaps in the record of hydrological processes4,5. Owing to its physicochemical properties, water responds clearly to changes in air temperature. This fact is widely used in studies of inland water thermal regimes, even in the absence of other climatic data. These studies often yield highly accurate results in explaining changes in surface water temperature across various temporal scales6,7. In the case of lakes, water temperature observations were already being conducted in the 19th century8. However, systematic measurements have only been regularly recorded since the early 20th century9. Therefore, lake temperature records do not have time series as long as those for air temperature, which for about thirty stations in Europe have been recorded continuously since the 18th century. Against this background, there is a clear lack of sufficient knowledge about the long-term thermal changes in lakes.At this point, it is important to highlight the fundamental role of water temperature in shaping the processes occurring in lakes10,11,12,13. In the long term, this influences the potential for using lakes for economic purposes such as irrigation, fisheries, tourism, and recreation. Today, a key issue is the response of lakes to climate change—an area of central importance in limnology— having detailed data allows for the interpretation of the magnitude of ecosystem changes occurring in lakes. In many regions around the world, a rise in surface water temperature has been observed, although the rate and scale of this warming vary14,15. Despite evidence of significant lake warming in recent decades, our understanding of long-term temperature changes remains relatively limited9. To fully understand the scope of the ongoing transformation in lake thermal characteristics over the past few decades, it is essential to collect information across various time scales. Expanding knowledge of lake temperature changes before the 20th century requires a reconstruction-based approach16,17.Meteorological reanalysis is becoming an increasingly common tool for data acquisition and is widely used in various studies related to hydrological issues18,19,20. In the case of water temperature, that was demonstrated for Lake Chaohu21, the use of reanalysis combined with hydrodynamic models can provide valuable insights into its dynamics.The main aim of the article is to reconstruct the annual and seasonal surface water temperature (LSWT) of selected lakes in Poland using the Twentieth Century Reanalysis (20CR) dataset. Based on the implementation of these assumptions, additional objectives were adopted, namely to determine the direction and magnitude of water temperature changes over the period 1806–2022. Achieving these objectives will provide an important starting point for further research on the thermal dynamics of inland waters, covering periods prior to significant human impact on the environment.Materials and methodsStudy areaThe study area covers lakes in the northern part of Poland (Fig. 1). The article analyses seven lakes, selected based on the availability of long-term surface water temperature measurements. All the lakes are of natural character and vary in morphometric parameters, with surface areas ranging from 2.44 to 70.20 km2 and mean depths from 1.6 to 11.6 m (Table 1). Notably, the geographical locations of the lakes place them under the influence of both maritime climate characteristics (western Poland) and continental climate features (eastern Poland). The mean air temperature ranges from 6.7 °C to 8.9 °C (east and west, respectively). In turn, the mean annual temperature of the analyzed lakes ranges from 9.3 °C (Łebsko) to 11 °C (Sławskie). The duration of the ice cover varies from 59 days (Lake Sławskie) to 96 days (Studzieniczne). Additionally, the northernmost Lake Łebsko is directly connected to the Baltic Sea, where one of the characteristics of coastal lakes is their shallow depth22. Cieśliński23, in determining the hydrochemical type of the water, points to dominant supply from chloride–sodium waters, indicating a constant influence of the Baltic Sea, where the average chloride concentration exceeds 750 mg dm⁻³.Fig. 1Studied lakes (figure generated in ArcGIS Pro v3.1.0 software), https://www.esri.com/en-us/arcgis/products/arcgis-pro/overview).Full size imageTable 1 Location and basic morphometric parameters of the analyzed lakes24*, 25**.Full size tableMaterialsTwo datasets were used in this study. The first one pertains to surface water temperature, obtained from measurements conducted by the Institute of Meteorology and Water Management – National Research Institute over the past 63 years (1960–2022). The data was available for all the lakes. Water temperature is routinely measured at a fixed point, always at the same location, at a depth of 0.4 m below the surface, at 6:00 UTC.The second one is the Twentieth-Century Reanalysis, version 3 (20CRv326). 20CRv3 is a comprehensive historical global reanalysis dataset that provides a range of atmospheric variables, including, among others, 2-meter above-ground-level (a.g.l.) air temperature. It covers the time period from 1806 to 2015 with the spatial resolution of 1.0 degree latitude x 1.0 degree longitude global grid (360 × 181). To achieve this extended temporal coverage, the reanalysis assimilates solely surface pressure observations. A detailed overview of the 20CR system, including a technical description of the data assimilation and the model used, is provided by Slivinski et al.26. 20CRv3 is capable of reliably generating atmospheric estimates across a range of scales, from individual weather events to long-term climate trends27. The monthly mean air temperature data at a height of 2 m a.g.l. utilised in this study were extracted from the 20CRv3 grid point nearest to the location of the studied lake (refer to Table 1), employing the nearest-neighbour remapping technique.Although 20CRv3 does not exhibit air temperature reconstruction biases for mid-latitudes when compared with other reanalyses for the contemporary period27, their Fig. 10], we nevertheless conducted an evaluation of the 1 × 1° gridded temperature data against long-term historical point measurements from Polish meteorological stations (Fig. S1), i.e. Gdańsk 1851–195928, Toruń 1871–195929, Warszawa 1806–195930. The available data range from the beginning of the measurements and/or the temporal coverage of 20CRv3 up to 1959, as this period of 20CRv3 gridded data was used as input for the reconstruction of lake water temperatures. The Pearson correlation (r) between the 20CRv3 gridded data and point measurements is very high (0.99) and statistically significant, with a coefficient of determination (R²) between 0.98 and 0.99, root mean square error (RMSE) of 0.73–1.16, and mean absolute error (MAE) of 0.55–0.86. Furthermore, when compared over the common period 1871–2015 for all stations (not shown), and with similarly high r and R² values, RMSE decreases to 0.63–0.73 and MAE to 0.48–0.54. Therefore, the use of 20CRv3 gridded data appears justified for reconstructing water temperature in lakes whose surface area is considerably smaller than that of a single grid cell.However, it should be kept in mind that the air temperature in 20CRv3 is underestimated for the period 1806–1850 (the mean annual difference between observational data from Warszawa and 20CRv3 is − 1.0 °C, see Fig. S2). This is due to the limited assimilation of input data into 20CRv3 prior to the year 185026, their Fig. 1].MethodsIn order to reconstruct the lake LSWT (Lake Surface Water Temperature) for the period 1806–1959, the air2water model was used. The air2water is a hybrid model that combines a physically-based equation (the surface layer energy balance) with stochastic calibration of the model parameters. Heat budget of the surface layer is calculated as follows:$$:rho:{C}_{p}Vfrac{dLSWT}{dt}={AH}_{net}$$
    (1)
    where: ρ – water density (1000 kg m− 3), Cp – specific heat capacity at a constant pressure (4186 J kg− 1 oC−1), V – surface volume (m3), LSWT – lake surface water temperature (oC), t – time in days, A – surface area (m2), Hnet – heat flux into the surface layer (W m− 2).The air2water model has been successfully used to study LSWT in various regions around the world7,31,32. In this study, the 6-parameter version of the air2water model was applied33.$$:frac{dLSWT}{dt}=frac{1}{delta:}left[{a}_{1}+{a}_{2}{T}_{air}-{a}_{3}LSWT+{a}_{c}text{c}text{o}text{s}left(2pi:left(frac{t}{{t}_{y}}-{a}_{6}right)right)right]$$
    (2)
    $$:delta:=left{begin{array}{cc}expleft(-frac{LSWT-{T}_{h}}{{a}_{4}}right)&:LSWTge:{T}_{h}\:1&:LSWT<{T}_{h}end{array}right.$$
    (3)
    where: Tair – air temperature (oC), a1, a2, a3, a4, a5 and a6 – model parameters determined during the process of model calibration and validation, ty – duration of a year (365 days), Th – reference value of the deep-water temperature (oC), δ – dimensionless term representing the ratio between the volume of the surface lake layer and a reference volume.To assess the usefulness of the air2water model, it was calibrated using data from the period 1960–1999 (40 years – approx. 63%), while data from the period 2000–2022 (23 years – approx. 27%) were used for model validation. Since this study uses monthly average water and air temperatures, the input data for the air2water model were prepared so that each day was assigned the corresponding monthly average air and water temperature value (Fig. 2). Based on this data, the model was calibrated. For the validation, the average monthly LSWT values were compared with the LSWT values from the 15th day of each month.Fig. 2Approach to modeling lake water temperatures using the air2water program – example results from 2022 for Lake Sławskie (AT – air temperature, LSWTp – water temperature obtained from the air2water model, LSWTo – water temperature based on measurement data).Full size imageTo assess the performance of the model, six commonly used metrics were used, including coefficient of determination (R2), root mean square error (RMSE), mean absolute error (MAE), the Nash-Sutcliffe efficiency coefficient (NSE) and the Kling-Gupta efficiency coefficient (KGE).To reconstruct the monthly average LSWT in the studied lakes for the period 1806–1959, the air2water model was calibrated individually for lakes. For this purpose, measurement data from the period 1960–2022 were used to determine the values of parameters a1, a2, a3, a4, a5, and a6. The values of R², RMSE, MAE, NSE, and KGE were recalculated and compared with the values obtained for the periods 1960–1999 (calibration) and 2000–2022 (validation using independent data).Based on the reconstructed values of monthly average water temperatures, annual and seasonal averages were calculated for spring (March – May), summer (Jun – August), autumn (September – November), and winter (December – February). The Pettitt test allows for the detection of single change points in time series. To enable its use for detecting multiple change points within the 1806–2022 period, 500 time series of lengths 30, 40, and 50 years were randomly selected without replacement, and the Pettitt test was performed for each series. This way, statistically significant breakpoints (at the 0.05 significance level) were identified, allowing for the detection of years in which changes were most likely to have occurred, individually for each lake. To standardize the analysis periods across all studied lakes, a regional change point was determined. A given year was defined as the regional change point if a change was detected in 4 out of the 7 lakes. If the change was identified in two consecutive years, the earlier year was selected as the regional change point. The analysis of long-term changes was conducted for the entire 1806–2022 period and for sub-periods. Additionally, it was assumed that the minimum length of any analysis period must be at least 20 years. For the analysis of long-term changes, the Mann-Kendall34 and Sen’s35 tests were used. The Mann-Kendall and Sen’s tests were performed using a modified version of the mk package developed by Patakamuri and O’Brien36. The detection of change points was carried out using the trend package developed by Pohlert37. Trend analysis and change point detection were conducted using the R software environment (Version R-4.5.2 (https://cran.r-project.org/).ResultsIn the first stage, the air2water model was calibrated and validated using monthly average air temperature (AT) and LSWT. During the calibration stage, the following results were obtained: R² ranging from 0.981 to 0.992, RMSE from 0.63 to 0.95 °C, MAE from 0.50 to 0.76 °C, NSE from 0.878 to 0.922, and KGE from 0.959 to 0.994. Lower goodness-of-fit metrics were obtained during the model validation stage: R² ranged from 0.980 to 0.995; NSE from 0.853 to 0.910; and KGE from 0.913 to 0.979, while RMSE values were higher, ranging from 0.68 to 1.20 °C, and MAE from 0.54 to 0.94 °C (Table 2). The model performance evaluation results suggest that the air2water model can be reliably used to reconstruct data for the period from 1806 to 1959.Table 2 The results of air2water model calibration (1960–1999) and validation (2000–2022).Full size tableDuring the reconstruction of lake water temperatures for the years 1806–1959, the air2water model was calibrated using all available measurement data from the period 1960 to 2022. The calibration quality results of the air2water model are presented in Table 3. The following model fit statistics were obtained: R² ranging from 0.973 to 0.981, RMSE from 0.77 to 1.15 °C, MAE from 0.60 to 0.90 °C, NSE from 0.857 to 0.900, and KGE from 0.972 to 0.992.Table 3 The results of air2water model calibration for the period of 1960–2022.Full size tableFig. 3presents the observed LSWT in comparison to those obtained from the air2water model.Full size imageFigure 3 Predicted vs. observed LSWT for lakes Lubie (a), Łebsko (b, ) Sławskie (c), Charzykowskie (d), Jeziorak (e), Nidzkie (f) i Studzieniczne (g). The black line represents the fit between observed and predicted LSWT, whereas the red represents the case if all predicted values perfectly matched the observed ones.The analysis of the annual average LSWT data series from 1806 to 2022 using Pettitt’s test revealed the presence of change points (Table S1). Between 9 and 12 change points were detected in the analyzed data series. In all lakes, changes were identified for the years 1987, 1998, and 2013; in six lakes for the years 1844 and 1988; and simultaneously in five lakes for the years 1909 and 1980. Pettitt’s test most frequently indicated a change point in the year 1844 (285 occurrences), followed by 1987 (191 occurrences). Additionally, in 1945, changes occurred 67 times across three lakes located in western Poland, and in 1988, changes were detected 88 times across six lakes.An analogous approach was applied for the spring, summer, autumn, and winter periods. The Pettitt test results revealed the presence of change points in different years (Tables S2–S5). Based on these results, so-called global change points were ultimately adopted, which allowed for the division of the data series into sub-series corresponding to the seasonal periods (Table S6). Based on the above assumptions, the analysis of changes in the average annual lake water temperatures was carried out using the Mann-Kendall and Sen’s tests for the periods 1806–1843 (38 years), 1844–1908 (65 years), 1909–1986 (78 years), and 1987–2022 (36 years) (Table 4). The average annual water temperatures in the lakes over the period 1806–2022 showed an increasing trend in all cases. The rate of temperature changes averaged 0.081 °C per decade, with the range of changes across individual cases varying from 0.049 to 1.05 °C per decade. Comparing values between two consecutive subperiods (determined based on Pettitt’s test analysis), an average temperature increase of 0.46 °C per decade was observed for 1806–1843 vs. 1844–1908; 0.40 °C per decade for 1844–1908 vs. 1909–1986; and as much as 0.81 °C per decade for 1909–1986 vs. 1987–2022 (Fig. 4).Table 4 Trend analysis results. Bold values mean statistically significant trends.Full size table Fig. 4Results of the annual mean LSWT trend analysis for lakes Lubie (a), Łebsko (b), Sławskie (c), Charzykowskie (d), Jeziorak (e), Nidzkie (f), and Studzieniczne (g) (solid lines represent significant trends at a level of 0.05). Additionally, the year-to-year course in air temperature of the grid point closest to the lake was plotted based on data from of the 20CRv326.Full size imageThe analysis of average spring water temperatures in lakes for the period 1806–2022 using Pettitt’s test revealed breakpoints in the data series for 17 years. The most frequent breakpoints occurred in 1845 (287 times), 1980 (141 times), and 1988 (132 times). The analysis of spring water temperature averages using the Mann-Kendall test showed an overall significant increase from 1806 to 2022 (mean value of 0.15 °C per decade). The analysis of average lake water temperatures during the summer period revealed breakpoints in 19 different years. The most frequent breakpoints occurred in 1991 — 200 times, in 1841 and 1987 — 98 times each, and in 1931 — 90 times. The analysis of average summer water temperatures using the Mann-Kendall test showed a significant increasing trend over the period 1806–2022 (average 0.08 °C per decade). The analysis of the average lake water temperatures during the autumn period from 1806 to 2022 using the Pettitt test showed that breakpoints were identified in the data series for 14 years. The most frequent breakpoints occurred in 1998–175 times, and in 1959–82 times. Average autumn water temperatures showed an increase in six lakes (0.07 °C per decade), while in Lake Łebsko a decrease in water temperature was observed (0.02 °C per decade). Changes in other periods were statistically insignificant. Analysis of average water temperatures in lakes during the winter period from 1806 to 2022 using the Pettitt test revealed breakpoints in the data series for as many as 21 years. The most frequent breakpoints occurred in 1897–166 times, in 1969–123 times, and in 1987–78 times. Average winter water temperatures increased in six lakes (significant at the 0.01 significance level) over the period 1806–2022 (0.05 °C per decade), while in Lake Łebsko a decrease in water temperature was observed (0.02 °C per decade).DiscussionUnderstanding the processes occurring in lakes relies on a diverse set of methodologies38,39,40,41, one of which is the reconstruction of historical conditions. Research on the reconstruction of water temperature is gaining increasing interest42,43,44, driven by the need to create a broader understanding of thermal changes in the hydrosphere45. Depending on the adopted methodology and reference data, such analyses cover different time intervals. In relation to lakes, it should be emphasized that a large portion of thermal studies focuses on the last several decades46,47,48, while less attention has been given to periods reaching back to the first half of the 20th century49,50. The use of reanalysis data combined with hydrodynamic models can provide information on the dynamics of water temperature in individual water bodies21. In line with these findings, the reconstruction of water temperature in seven lakes in Poland presented in this article significantly expands the current knowledge on inland water thermal dynamics, covering a period of over 200 years.Considering the changes in water temperature in the studied lakes, several distinct phases can be observed, which generally reflect shifts in climatic conditions. Based on the Pettitt test results, characteristic moments include the 1840s, the 1940s, and the late 1980s. The beginning of the analysed period is marked by the lowest temperatures and a decreasing trend. The first decades of the 19th century (up to 1840) in Poland were characterized by a greater degree of thermal continentality than that observed today. In a broader perspective, this represented the final phase of a climatic situation that had persisted since the 16th century51. From this point onward, water temperatures gradually increased, although some downward tendencies can be noted. One such period occurred from the mid-1940s, following one of the greatest warming phases of the 20th century, during which global temperatures rose by 0.37 °C between 1925 and 194452. According to the Sen’s test, the key moment of thermal regime transformation occurred the late 1980s, when a marked warming of lake waters occurred relative to the preceding period. This change corresponds to a shift in the climatic regime, which also influenced lake temperatures. Similar observations were confirmed by previous studies conducted on 20 lakes in Central Europe53, where six lakes from the area of Poland were included. Furthermore, in two additional lakes (Studzieniczne and Białe Augustowskie, northeastern Poland), it was determined that a significant change occurred at the same time54.The results obtained in this study are consistent with other research on water temperature reconstructions. For example, over the past 150 years of inland water monitoring (Pannonian ecoregion, Europe), a clear warming trend has been observed, with most of the warming occurring in recent decades55. Significant changes over the last few decades are also evident in many other cases—for instance, the increase in Vrana Lake’s water temperature (Croatia) was particularly pronounced after 201356. Trend analysis for Lake Miedwie (northwestern Poland) showed an average warming rate of 0.20 °C/decade, with the last thirty years of this period exhibiting an accelerated increase of 0.31 °C/decade57.Throughout the entire analysis period from 1806 to 2022, the rate of change in water temperature varied widely, ranging from 0.049 °C per decade (Sławskie Lake) to 0.105 °C per decade (Studzieniczne Lake). Considering the extreme locations of these two cases—southwest and northeast Poland respectively—this variation should be explained by the characteristics of the regions where they are located. The northeastern part of Poland is characterized by a continental climate, one of whose distinctive features is colder and longer winters compared to the west. In relation to the hydrosphere, this translates into the duration of ice cover phenomena. This situation is changing with increasing global warming, resulting in a later onset of ice formation and an earlier ice break-up date. According to data collected for the period 1961–2010, the average duration of ice cover on Lake Sławskie was 59.2 days, whereas on Lake Studzieniczne it was over a month longer (96.4 days)58. Until recently, Lake Studzieniczne was effectively isolated from external (atmospheric) influences for one quarter of the year. The recorded changes in ice cover duration indicate that the rate of ice cover decreased by an average of 3.7 days per decade in the first case, and as much as 6.1 days per decade in the second58. Consequently, the increasingly shorter ice season leads to longer periods of water warming in lakes, which is reflected in a higher rate of increase in water temperature. The earlier onset of the stratification season in lakes was significant for heat storage and average surface water temperature59. Seasonal change analysis showed the highest increase during spring (0.15 °C/decade), which can also be attributed to the earlier ice cover break-up dates. Even in the 1960s, in many Polish lakes the average ice disappearance date fell in the third decade of March, while today it is at the end of February60. Furthermore, against the backdrop of seasonal data, a different response was recorded in Lake Łebsko to water temperature changes was observed compared to the other lakes. This situation is caused by two factors: depth and location. Lake Łebsko is a polymictic lake, similar to Lake Sławskie, in which, however, such seasonal reactions were not observed. Cieśliński and Chlost61 indicate that factors potentially influencing water temperature include the intensity of water exchange and the magnitude of marine water intrusions.Considering the fundamental importance of water temperature for inland waters, the results obtained in this study should be regarded as unfavorable. The overall direction of changes observed over more than 200 years is unequivocal, indicating a permanent warming trend. This transformation poses a threat to lakes, causing disturbances in their functional balance. Increasing warming of the surface water layer will affect the stability of the water column. Yang et al.62 indicate that the development of thermal stratification is an important factor regulating the composition and abundance of phytoplankton during the summer period. As noted by Oleksy and Richardson63, an increase in the intensity and duration of stratification in dimictic lakes can alter the mixing regime of monomictic lakes, resulting in oxygen deficits in the hypolimnion, as well as changes in biogeochemistry and productivity. This concerns, among others, water quality issues, which is confirmed by studies such as those on Lake Yangzong (China), where water quality parameters were shown to be significantly correlated with and dependent on temperature64. Polish regulations concerning the classification and assessment of surface water bodies refer to the legal acts (directives) of the European Union, according to which the general status of the analyzed lakes is bad. Furthermore, the catchment areas of these lakes are sensitive to eutrophication, which leads, among other effects, to accelerated algae growth. Ongoing climate changes will increase the risk of cyanobacterial blooms in northern lakes, where in subarctic Quebec (Canada) the cyanobacterial community biovolumes positively correlated with surface water temperatures65. Similar situations have been observed in other regions; for example, in China, warming of Lake Dianchi’s surface temperature has extended the risk period for algal blooms and showed a positive correlation with algal density66. It should also be noted that water quality itself can influence water temperature. Previous analyses23 including the lakes currently under study, referred to these relationships by considering water transparency. PCA analysis showed negative relationships, where a decrease in transparency leads to an increase in water temperature due to greater absorption of solar radiation in the surface water layer.In the context of the relationship between water temperature and its quality, oxygen concentration is a key factor, because it decreases with rising temperature. This limits the water’s self-purification capacity and affects aerobic organisms67. Many factors influence fish distribution, with water temperature and dissolved oxygen being particularly restrictive68. Research on Lake Tanganyika showed that climate warming and intensified stratification reduced the lake’s potential fish production, leading to decreased fish catches69. Large and deep lakes will likely serve as thermal refuges for cold- and cool-water fish species even as average lake temperatures rise70. Among the seven lakes analyzed, three have average depths not exceeding 6 m, and it is in these shallower lakes that the fastest changes in ichthyofauna composition due to rising temperatures are expected. Changes in thermal thresholds will be crucial for hydrobiological shifts. Potential gains in species numbers from warmer waters may not fully compensate for losses of cold-water species with ongoing warming71. Previous studies of inland water ichthyofauna in northern Poland indicate that species with upper thermal tolerance limits below 28 °C live at the edge of their range72. The observed changes over the past two centuries allow us to conclude that the last few decades are particularly alarming, with a marked increase in water temperature. According to current climate scenarios, the trends observed in recent years are expected to continue54. The scale of ongoing and anticipated future changes necessitates actions to mitigate the effects of lake ecosystem transformation.ConclusionIn the case of the hydrosphere, water temperature is one of its key parameters, with the distribution and changes in temperature influencing the functioning and transformation of its individual components. This article presents an analysis of the thermal regime of seven lakes in Central Europe over an unprecedented period spanning 1806–2022. The use of the air2water model, which utilizes air temperature data from the 20CRv3 reanalysis, proved to be an effective approach, as confirmed by high statistical test results. Overall, the observed changes reflect the prevailing climatic conditions. Across all cases, an increase in the average annual water temperature of 0.081 °C per decade in the period of 1806–2022 was recorded, with individual lakes exhibiting rates ranging from 0.049 to 0.105 °C per decade. Notably, the most significant increases were observed over the last few decades, and current studies suggest this warming trend will continue in the future. The results obtained in this study are unfavorable with regard to lake functioning, as the more than two-century-long warming will drive their transformation, contributing to declines in water quality and alterations in hydrobiological conditions. This situation calls for multidisciplinary consultations and subsequent actions aimed at developing strategies to mitigate the impacts of global warming on lake ecosystems.

    Data availability

    Datasets for this research were derived from the following public domain resources:- Lake Surface Water Temperature: Institute of Meteorology and Water Management – National Research Institute (IMGW-PIB) [https://danepubliczne.imgw.pl/data/dane_pomiarowo_obserwacyjne/dane_hydrologiczne/] for the period 1984-2022 and data transcribed from Hydrological Yearbooks of IMGW-PIB, 1960-1983 by the first Author which are available on reasonable request.- Air Temperature: 20th Century Reanalysis (V3), [https://psl.noaa.gov/data/gridded/data.20thC_ReanV3.html] . These data were produced by National Oceanic and Atmospheric Administration (NOAA) and are not subject to copyright protection in the United States. NOAA waives any potential copyright and related rights in these data worldwide through the Creative Commons Zero 1.0 Universal Public Domain Dedication (CC0-1.0).
    ReferencesKolář, P. Changes and variability of spring-summer air temperature in Czechia during the past 300 years: comparison of instrumental, documentary and natural proxy data. Geografie-Sbornik CGS. 122 (2), 190–212. https://doi.org/10.37040/geografie2017122020190 (2017).Article 

    Google Scholar 
    Bravo-Paredes, N., Gallego, M. C., Trigo, R. M. & Vaquero, J. M. Earliest meteorological readings in San Fernando (Cádiz, Spain, 1799–1813). Clim. Past 19(7), 1397–1408 (2023). https://doi.org/10.5194/cp-19-1397-2023Kolendowicz, L. et al. Homogenization of air temperature and its long-term trends in Poznań (Poland) for the period 1848–2016. Theoret. Appl. Climatol. 136, 1357–1370. https://doi.org/10.1007/s00704-018-2560-z (2019).Article 

    Google Scholar 
    Llasat, M. C., Barriendos, M., Barrera, A. & Rigo, T. Floods in Catalonia (NE Spain) since the 14th century. Climatological and meteorological aspects from historical documentary sources and old instrumental records. J. Hydrol. 313 (1–2), 32–47. https://doi.org/10.1016/j.jhydrol.2005.02.004 (2005).Article 

    Google Scholar 
    Schoning, K., Charman, D. J. & Wastegård, S. Reconstructed water tables from two ombrotrophic mires in Eastern central Sweden compared with instrumental meteorological data. Holocene 15 (1), 111–118. https://doi.org/10.1191/0959683605hl772rp (2005).Article 

    Google Scholar 
    Naumenko, M. A. & Guzivaty, V. V. Climate relationships between air temperature and water temperatures in the different limnetic zones of lake Ladoga. Geogr. Nat. Resour. 43 (1), 69–76. https://doi.org/10.1134/S1875372822010097 (2022).Article 

    Google Scholar 
    Heddam, S., Ptak, M. & Zhu, S. Modelling of daily lake surface water temperature from air temperature: extremely randomized trees (ERT) versus air2Water, MARS, M5Tree, RF and MLPNN. J. Hydrol. 588, 125130. https://doi.org/10.1016/j.jhydrol.2020.125130 (2020).Article 

    Google Scholar 
    Świerz, L. Pomiary ciepłoty stawów tatrzańskich w różnych Warstwach głębokości. Pamiętnik Towarzystwa Tatrzańskiego 10, 122 (1885).Piccolroaz, S., Woolway, R. I. & Merchant, C. J. Global reconstruction of twentieth century lake surface water temperature reveals different warming trends depending on the climatic zone. Clim. Chang. 160, 427–442. https://doi.org/10.1007/s10584-020-02663-z (2020).Article 

    Google Scholar 
    Barrett, P. M. et al. Contrasting arsenic cycling in strongly and weakly stratified contaminated lakes: evidence for temperature control on sediment–water arsenic fluxes. Limnol. Oceanogr. 64 (3), 1333–1346. https://doi.org/10.1002/lno.11119 (2019).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Li, Y., Xia, M., Zhao, X. & Hou, H. Water temperature and chlorophyll a density drive the genetic and epigenetic variation of vallisneria natans across a subtropical freshwater lake. Ecol. Evol. 13 (8), e10434. https://doi.org/10.1002/ece3.10434 (2023).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Nakayama, K., Komai, K., Amano, M., Kumamoto, E. & Oyama, Y. Ideal water temperature environment for giant Marimo (Aegagropila linnaei) in Lake Akan, Japan. Sci. Rep. 13 (1), 16834. https://doi.org/10.1038/s41598-023-43792-6 (2023).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Fukushima, T., Setiawan, F., Subehi, L., Jiang, D. & Matsushita, B. Water temperature and some water quality in lake Toba, a tropical volcanic lake. Limnology 24 (1), 61–69. https://doi.org/10.1007/s10201-022-00703-4 (2023).Article 
    CAS 

    Google Scholar 
    Haddout, S., Qanza, H., Guennoun, M. A., Karra, R. & Essaidi, A. Epilimnion and Metalimnion thermal water temperature variables in moroccan’s lakes using a one-dimensional fresh-water lake model. Int. J. River Basin Manage. 18 (3), 321–333. https://doi.org/10.1080/15715124.2018.1498854 (2020).Article 

    Google Scholar 
    Virdis, S. G. P., Soodcharoen, N., Lugliè, A. & Padedda, B. M. Estimation of satellite-derived lake water surface temperatures in the Western mediterranean: integrating multi-source, multi-resolution imagery and a long-term field dataset using a time series approach. Sci. Total Environ. 707, 135567. https://doi.org/10.1016/j.scitotenv.2019.135567 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Tierney, J. E. et al. Late-twentieth-century warming in lake Tanganyika unprecedented since AD 500. Nat. Geosci. 3 (6), 422–425. https://doi.org/10.1038/ngeo865 (2010).Article 
    CAS 

    Google Scholar 
    Matulla, C. et al. Establishment of a long-term lake-surface temperature dataset within the European alps extending back to 1880. Clim. Dyn. 52 (9–10), 5673–5689. https://doi.org/10.1007/s00382-018-4479-6 (2019).Article 

    Google Scholar 
    Sabarly, F., Essou, G., Lucas-Picher, P., Poulin, A. & Brissette, F. Use of four reanalysis datasets to assess the terrestrial branch of the water cycle over Quebec, Canada. J. Hydrometeorol. 17 (5), 1447–1466. https://doi.org/10.1175/JHM-D-15-0093.1 (2016).Article 

    Google Scholar 
    Mahto, S. S., Mishra, V., Does, V. & 5 outperform other reanalysis products for hydrologic applications in India? J. Geophys. Research: Atmos. 124 (16), 9423–9441. https://doi.org/10.1029/2019JD031155 (2019).Article 

    Google Scholar 
    Singh, H. & Mohanty, M. P. Can atmospheric reanalysis datasets reproduce flood inundation at regional scales? A systematic analysis with ERA5 over Mahanadi River Basin, India. Environ. Monit. Assess. 195 (10), 1143. https://doi.org/10.1007/s10661-023-11798-2 (2023).Article 
    PubMed 

    Google Scholar 
    Frassl, M. A. et al. Opportunities and limits of using meteorological reanalysis data for simulating seasonal to sub-daily water temperature dynamics in a large shallow lake. Water 10, 594 (2018). https://doi.org/10.3390/w10050594Choiński, A., Ptak, M. & Strzelczak, A. Present-day evolution of coastal lakes based on the example of Jamno and Bukowo (the Southern Baltic coast). Oceanological Hydrobiol. Stud. 43 (2), 178–184. https://doi.org/10.2478/s13545-014-0131-1 (2014).Article 

    Google Scholar 
    Cieśliński, R. Zróżnicowanie typologiczne i Funkcjonalne Jezior w Polskiej strefie Brzegowej południowego Bałtyku. Probl. Ekologii Krajobrazu 26, 135–144 (2010).Choiński, A. Katalog Jezior Polski (Wyd. Nauk UAM, 2006).Ptak, M., Sojka, M., Choiński, A. & Nowak, B. Effect of environmental conditions and morphometric parameters on surface water temperature in Polish lakes. Water 10, 580. https://doi.org/10.3390/w10050580 (2018).Article 

    Google Scholar 
    Slivinski, L. C. et al. Towards a more reliable historical reanalysis: improvements for version 3 of the twentieth century reanalysis system. Q. J. R Meteorol. Soc. 145, 2876–2908. https://doi.org/10.1002/qj.3598 (2019).Article 

    Google Scholar 
    Slivinski, L. C. An evaluation of the performance of the 20th century reanalysis version 3. J. Clim. 34 (4), 1417–1438. https://doi.org/10.1175/JCLI-D-20-0505.1 (2021).Article 

    Google Scholar 
    Miętus, M. The reconstruction and homogenization of long-term series of monthly mean temperature from Gdańsk–Wrzeszcz station, 1851–1995. Wiadomości IMGW. 21, 41–63 (1998).
    Google Scholar 
    Pospieszyńska, A. & Przybylak, R. Air temperature changes in Toruń (central Poland) from 1871 to 2010. Theor. Appl. Climatol. 135, 707–724. https://doi.org/10.1007/s00704-018-2413-( (2019).Article 

    Google Scholar 
    Lorenc, H. Studia Nad 220-letnią (1779–1998) serią temperatury Powietrza w Warszawie Oraz Ocena Jej Wiekowych Tendencji. Mat. Bad Ser. Meteorologia. 31, 3–1 (2000).
    Google Scholar 
    Piccolroaz, S. et al. On the predictability of lake surface temperature using air temperature in a changing climate: a case study for lake Tahoe (U.S.A). Limnol. Oceanogr. 63, 243–261. https://doi.org/10.1002/lno.10626 (2018).Article 

    Google Scholar 
    Jia, T. & Luo, Y. Using Air2water model to predict the water temperature of fuxian lake. In Climate Change and Environmental Sustainability, Advances in Science, Technology & Innovation (eds He, B. J., Prasad, D., Pignatta, G. & Jupesta, J.). 29–35 . https://doi.org/10.1007/978-3-031-12015-2_4 (Springer, 2022).Chapter 

    Google Scholar 
    Piccolroaz, S. Prediction of lake surface temperature using the air2water model: guidelines, challenges, and future perspectives. Adv. Oceanogr. Limnol. 7 (1), 36–50. https://doi.org/10.4081/aiol.2016.5791 (2016).Article 

    Google Scholar 
    Kendall, M. G. & Stuart, A. The Advanced Theory of Statistics. 3rd Ed. (Charles Griffin Ltd., 1986).Gilbert, R. O. Statistical Methods for Environmental Pollution Monitoring (Van Nos-trand Reinhold Co., 1987).Patakamuri, S. K. & O’Brien, N. Modified Versions of Mann Kendall and Spearman’s Rho Trend Tests, Version 1.6. (2022). https://CRAN.R-project.org/package=modifiedmkPohlert, T. Non-Parametric trend tests and Change-Point detection, Version 1.1.6. (2023). https://CRAN.R-project.org/package=trendPtak, M., Choiński, A., Strzelczak, A. & Targosz, A. Disappearance of Lake Jelenino since the end of the XVIII century as an effect of anthropogenic transformations of the natural environment. Pol. J. Environ. Stud. 22 (1), 191–196 (2013).
    Google Scholar 
    Choiński, A., Ptak, M. & Strzelczak, A. Changeability of accumulated heat content in alpine-type lakes. Pol. J. Environ. Stud. 24 (6), 2363–2369. https://doi.org/10.15244/pjoes/58871 (2015).Article 

    Google Scholar 
    Tylmann, W. et al. Investigating limnological processes and modern sedimentation at lake Żabińskie, Northeast Poland: a decade-long multi-variable dataset, 2012–2021. Earth Syst. Sci. Data. 15, 5093–5103. https://doi.org/10.5194/essd-15-5093-2023 (2023).Article 

    Google Scholar 
    Zawiska, I., Woszczyk, M. & Rzodkiewicz, M. Towards a quantitative reconstruction of lake trophic state in temperate lakes using subfossil cladocera and diatoms: composition of a training set from NE Poland. Geographia Pol. 95 (3), 227–253. https://doi.org/10.7163/GPol.0234 (2022).Article 

    Google Scholar 
    Livingstone, D. M. & Lotter, A. F. The relationship between air and water temperatures in lakes of the Swiss Plateau: A case study with Palaeolimnological implications. J. Paleolimnol. 19 (2), 181–198. https://doi.org/10.1023/A:1007904817619 (1998).Article 

    Google Scholar 
    Zhang, X., Wang, K., Frassl, M. A. & Boehrer, B. Reconstructing six decades of surface temperatures at a shallow lake. Water 12, 405. https://doi.org/10.3390/w12020405 (2020).Article 

    Google Scholar 
    Shrestha, R. R. & Pesklevits, J. C. Reconstructed River Water Temperature Dataset for Western Canada 1980–2018. Data 8, 48. https://doi.org/10.3390/data8030048 (2023).Sojka, M. & Ptak, M. Possibilities of river water temperature reconstruction using statistical models in the context of long-term thermal regime changes assessment. Appl. Sci. 12, 7503, 1–18. https://doi.org/10.3390/app12157503 (2022).Article 
    CAS 

    Google Scholar 
    Haddout, S., Priya, K. & Boko, M. Thermal response of Moroccan lakes to climatic warming: first results. Ann. Limnol. Int. J. Limnol. 54, 2. https://doi.org/10.1051/limn/2017029 (2018).Article 

    Google Scholar 
    Yang, K., Yu, Z., Luo, Y., Zhou, X. & Shang, C. Spatial-temporal variation of lake surface water temperature and its driving factors in Yunnan-Guizhou Plateau. Water Resour. Res. 55, 4688–4703. https://doi.org/10.1029/2019WR025316 (2019).Article 

    Google Scholar 
    Peng, Z. et al. Attribution analysis of lake surface water temperature changing—Taking China’s six main lakes as example. Ecol. Indic. 145, 109651. https://doi.org/10.1016/j.ecolind.2022.109651 (2022).Article 

    Google Scholar 
    Austin, J. & Colman, S. A century of temperature variability in lake superior. Limnol. Oceanogr. 53, 2724–2730. https://doi.org/10.4319/lo.2008.53.6.2724 (2008).Article 

    Google Scholar 
    Magee, M. R., Wu, C. H., Robertson, D. M., Lathrop, R. C. & Hamilton, D. P. Trends and abrupt changes in 104 years of ice cover and water temperature in a dimictic lake in response to air temperature, wind speed, and water clarity drivers. Hydrol. Earth Syst. Sci. 20, 1681–1702. https://doi.org/10.5194/hess-20-1681-2016 (2016).Article 

    Google Scholar 
    Przybylak, R. Zmiany Klimatu w Polsce i Europie w Ostatnich Stuleciach. Kosm Probl. Nauk. Biol. 57, 195–208 (2008).
    Google Scholar 
    Jones, P., New, M., Parker, D. E., Martin, S. & Rigor, I. G. Surface air temperature and its changes over the past 150 years. Rev. Geophys. 37, 173–199. https://doi.org/10.1029/1999RG900002 (1999).Article 

    Google Scholar 
    Woolway, R. I., Dokulil, M. T., Marszelewski, W., Schmid, M. & Bouffard, D. Warming of Central European lakes and their response to the 1980s climate regime shift. Clim. Change. 42, 505–520. https://doi.org/10.1007/s10584-017-1966-4 (2017).Article 

    Google Scholar 
    Ptak, M., Amnuaylojaroen, T. & Sojka, M. Seven decades of surface temperature changes in central European lakes. What’s next? Resources. 13 (149), 1–16. https://doi.org/10.3390/resources13110149 (2024).Article 

    Google Scholar 
    Li, H. et al. 150-year daily data (1870 – 2021) in lakes and rivers unveils escalating surface water warming and intensifying heatwaves in the Pannonian ecoregion (Hungary). J. Hydrology: Reg. Stud. 56, 101985, 1–15. https://doi.org/10.1016/j.ejrh.2024.101985 (2024).Article 

    Google Scholar 
    Brönnimann, S. & Brugnara, Y. The weather diaries of the Kirch family: Leipzig, Guben, and Berlin (1677–1774). Clim. Past 19(7), 1435–1445 (2023). https://doi.org/10.5194/cp-19-1435-2023Ptak, M. et al. Utilizing multi-source datasets for the reconstruction and prediction of water temperature in lake Miedwie (Poland). Remote Sens. 16, 2753, 1–17. https://doi.org/10.3390/rs16152753 (2024).Article 

    Google Scholar 
    Choiński, A., Ptak, M., Skowron, R. & Strzelczak, A. Changes in ice phenology on Polish lakes from 1961–2010 related to location and morphometry. Limnologica 53, 42–49. https://doi.org/10.1016/j.limno.2015.05.005 (2015).Article 

    Google Scholar 
    Mishra, V., Cherkauer, K. A. & Bowling, L. C. Changing thermal dynamics of lakes in the great lakes region: role of ice cover feedbacks. Glob Planet. Chang. 75, 155–172. https://doi.org/10.1016/j.gloplacha.2010.11.003 (2011).Article 

    Google Scholar 
    Ptak, M. & Sojka, M. The disappearance of ice cover on temperate lakes (Central Europe) as a result of global warming. Geographical J. 187 (3), 200–213. https://doi.org/10.1111/geoj.12385 (2021).Article 

    Google Scholar 
    Cieśliński, R. & Chlost, I. Charakterystyka Wybranych Cech Hydrologicznych Jezior Przymorskich Jako Konsekwencja Zmiany Klimatu. Przegląd Geograficzny 96(3) (2024).Yang, Y., Colom, W., Pierson, D. & Pettersson, K. Water column stability and summer phytoplankton dynamics in a temperate lake (Lake Erken, Sweden). Inland. Waters. 6 (4), 499–508. https://doi.org/10.1080/IW-6.4.874 (2016).Article 
    CAS 

    Google Scholar 
    Oleksy, I. A. & Richardson, D. C. Climate change and teleconnections amplify lake stratification with differential local controls of surface water warming and deep water cooling. Geophys. Res. Lett. 48 (5), e2020GL090959. https://doi.org/10.1029/2020GL090959 (2021).Xu, W. et al. Effects of seasonal variation on water quality parameters and eutrophication in Lake Yangzong. Water 14 (17), 2732. https://doi.org/10.3390/w14172732 (2022).Article 
    CAS 

    Google Scholar 
    Przytulska, A., Bartosiewicz, M. & Vincent, W. F. Increased risk of cyanobacterial blooms in northern high-latitude lakes through climate warming and phosphorus enrichment. Freshw. Biol. 62 (12), 1986–1996. https://doi.org/10.1111/fwb.13043 (2017).Article 
    CAS 

    Google Scholar 
    Peng, J. et al. Dynamics of the risk of algal blooms induced by surface water temperature in an alpine eutrophic lake under climate warming: insights from lake dianchi. J. Hydrol. 643, 131949. https://doi.org/10.1016/j.jhydrol.2024.131949 (2024).Article 

    Google Scholar 
    Ptak, M. & Nowak, B. Variability of oxygen-thermal conditions in selected lakes in Poland. Ecol. Chem. Eng. S. 23 (4), 639–650. https://doi.org/10.1515/eces-2016-0045 (2016).Article 
    CAS 

    Google Scholar 
    Brooks, J. L. et al. The influence of thermal and hypoxia induced habitat compression on Walleye (Sander vitreus) movements in a temperate lake. Mov. Ecol. 13, 1. https://doi.org/10.1186/s40462-024-00505-6 (2025).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Cohen, A. S. et al. Climate warming reduces fish production and benthic habitat in Lake Tanganyika, one of the most biodiverse freshwater ecosystems. Proc. Natl. Acad. Sci. U.S.A. 113 (34), 9563–9568. https://doi.org/10.1073/pnas.1603237113 (2016).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wszola, L. et al. Lake temperature and morphometry shape the thermal composition of recreational fishing catch. Trans. Am. Fish. Soc. 153 (6), 746–762. https://doi.org/10.1002/tafs.10481 (2024).Article 
    CAS 

    Google Scholar 
    Xu, L. et al. Asymmetric impacts of climate change on thermal habitat suitability for inland lake fishes. Nat. Commun. 1, 10273. https://doi.org/10.1038/s41467-024-54533-2 (2024).Article 
    CAS 

    Google Scholar 
    Radtke, G. & Bernaś, R. Temperature tolerance of European fish species based on thermal maxima in southern Baltic Sea-basin streams. Ecol. Ind. 170, 113107. https://doi.org/10.1016/j.ecolind.2025.113107 (2025).Article 

    Google Scholar 
    Download referencesAcknowledgementsSupport for the Twentieth Century Reanalysis Project version 3 dataset is provided by the U.S. Department of Energy, Office of Science Biological and Environmental Research (BER), by the National Oceanic and Atmospheric Administration Climate Program Office, and by the NOAA Earth System Research Laboratory Physical Sciences Laboratory.FundingR.P. and P.W. have been supported by the National Science Centre, Poland (grant no. 2020/37/B/ST10/00710).Author informationAuthors and AffiliationsDepartment of Hydrology and Water Management, Adam Mickiewicz University, Krygowskiego 10, 61-680, Poznań, PolandMariusz PtakFaculty of Earth Sciences and Spatial Management, Nicolaus Copernicus University, Lwowska 1, 87-100, Toruń, PolandRajmund Przybylak & Przemysław WyszyńskiCentre for Climate Change Research, Nicolaus Copernicus University, Lwowska 1, 87-100, Toruń, PolandRajmund Przybylak & Przemysław WyszyńskiDepartment of Land Improvement, Environmental Development and Spatial Management, Poznań University of Life Sciences, Poznań, Piątkowska 94E, 60-649, Poznań, PolandMariusz SojkaAuthorsMariusz PtakView author publicationsSearch author on:PubMed Google ScholarRajmund PrzybylakView author publicationsSearch author on:PubMed Google ScholarPrzemysław WyszyńskiView author publicationsSearch author on:PubMed Google ScholarMariusz SojkaView author publicationsSearch author on:PubMed Google ScholarContributionsM.P. : Conceptualization, Investigation, Data collection and selection, Interpretation of results, Writing – original draft, Writing – review & editing, Project administration. R.P. : Conceptualization, Methodology, Data collection and selection, Writing-original draft, Writing – review & editing. P.W. : Conceptualization, Methodology, Data collection and selection, Data curation, Software, Validation, Writing – original draft, Writing – review & editing. M.S. : Conceptualization, Methodology, Investigation, Data curation, Interpretation of results, Writing – original draft, Writing – review & editing, Software, Visualisation.Corresponding authorCorrespondence to
    Przemysław Wyszyński.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Additional informationPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary InformationBelow is the link to the electronic supplementary material.Supplementary Material 1Rights and permissions
    Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
    Reprints and permissionsAbout this articleCite this articlePtak, M., Przybylak, R., Wyszyński, P. et al. Twentieth Century Reanalysis version 3 as a source of information on long-term trends (1806–2022) in lake surface water temperature changes in Central Europe (Poland).
    Sci Rep 15, 43833 (2025). https://doi.org/10.1038/s41598-025-28581-7Download citationReceived: 23 July 2025Accepted: 11 November 2025Published: 15 December 2025Version of record: 15 December 2025DOI: https://doi.org/10.1038/s41598-025-28581-7Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative
    KeywordsWater temperatureClimatic changeReconstructionTrendPoland More

  • in

    Advancing an adaptable and practical framework to address water quality challenges in a changing world

    As water-quality challenges intensify, the widely used Weighted Regressions on Time, Discharge, and Season (WRTDS) method offers an adaptable and practical framework for global water-quality science and management.

    Access through your institution

    Buy or subscribe

    This is a preview of subscription content, access via your institution

    Access options

    Access through your institution

    Access Nature and 54 other Nature Portfolio journals

    Get Nature+, our best-value online-access subscription

    $32.99 / 30 days

    cancel any time

    Learn more

    Subscribe to this journal

    Receive 12 digital issues and online access to articles

    $119.00 per year
    only $9.92 per issue

    Learn more

    Buy this articlePurchase on SpringerLinkInstant access to the full article PDF.USD 39.95Prices may be subject to local taxes which are calculated during checkout

    Additional access options:

    Log in

    Learn about institutional subscriptions

    Read our FAQs

    Contact customer support

    ReferencesHelsel, D. R. et al. Statistical Methods in Water Resources (USGS, 2020); https://doi.org/10.3133/tm4a3.Hirsch, R. M. et al. Weighted Regressions on Time, Discharge, and Season (WRTDS), with an application to Chesapeake Bay river inputs. J. Am. Water Resour. Assoc. 46, 857–880 (2010).Article 

    Google Scholar 
    Hirsch, R. M. et al. A bootstrap method for estimating uncertainty of water quality trends. Environ. Model. Softw. 73, 148–166 (2015).Article 

    Google Scholar 
    Choquette, A. F. et al. Tracking changes in nutrient delivery to western Lake Erie: Approaches to compensate for variability and trends in streamflow. J. Great Lakes Res. 45, 21–39 (2019).Article 
    CAS 

    Google Scholar 
    Zhang, Q. & Hirsch, R. M. River water‐quality concentration and flux estimation can be improved by accounting for serial correlation through an autoregressive model. Water Resour. Res. 55, 9705–9723 (2019).Article 

    Google Scholar 
    Pellerin, B. A. et al. Mississippi River nitrate loads from high frequency sensor measurements and regression-based load estimation. Environ. Sci. Technol. 48, 12612–12619 (2014).Article 
    CAS 

    Google Scholar 
    DeCicco, L. A. et al. WRTDSplus: Extensions to the WRTDS Method (USGS, 2024); https://doi.org/10.5066/P14PE5AN.Green, C. T. et al. Projecting stream water quality using Weighted Regression on Time, Discharge, and Season (WRTDS): An example with drought conditions in the Delaware River Basin. Sci. Total Environ. 999, 180286 (2025).Article 
    CAS 

    Google Scholar 
    Vecchia, A. V. et al. Modeling variability and trends in pesticide concentrations in streams. J. Am. Water Resour. Assoc. 44, 1308–1324 (2008).Article 

    Google Scholar 
    Fang, K. et al. Modeling continental US stream water quality using long-short term memory and weighted regressions on time, discharge, and season. Front. Water 6, 1456647 (2024).Article 

    Google Scholar 
    Download referencesAcknowledgementsThe research of the authors is supported by funding from the US Environmental Protection Agency and the US Geological Survey. The authors thank the broader community of researchers who have applied, tested and advanced WRTDS over the past 15 years. Any use of trade, firm, or product names is for descriptive purposes only and does not imply endorsement by the US Government. This is UMCES contribution number 6481.Author informationAuthors and AffiliationsUniversity of Maryland Center for Environmental Science, Annapolis, MD, USAQian ZhangUS Environmental Protection Agency Chesapeake Bay Program, Annapolis, MD, USAQian ZhangUS Geological Survey, Reston, VA, USARobert M. HirschUS Geological Survey, Madison, WA, USALaura A. DeCiccoUS Geological Survey, DeKalb, IL, USAJennifer C. MurphyAuthorsQian ZhangView author publicationsSearch author on:PubMed Google ScholarRobert M. HirschView author publicationsSearch author on:PubMed Google ScholarLaura A. DeCiccoView author publicationsSearch author on:PubMed Google ScholarJennifer C. MurphyView author publicationsSearch author on:PubMed Google ScholarCorresponding authorCorrespondence to
    Qian Zhang.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Additional informationRelated linksEGRET webpage: https://rconnect.usgs.gov/EGRET/Rights and permissionsReprints and permissionsAbout this articleCite this articleZhang, Q., Hirsch, R.M., DeCicco, L.A. et al. Advancing an adaptable and practical framework to address water quality challenges in a changing world.
    Nat Rev Earth Environ (2025). https://doi.org/10.1038/s43017-025-00753-zDownload citationPublished: 15 December 2025Version of record: 15 December 2025DOI: https://doi.org/10.1038/s43017-025-00753-zShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Unlocking global rainwater harvesting potential for safe drinking water access

    AbstractSustainable Development Goal (SDG) 6.1 aims for universal access to safely managed drinking water (SMDW) by 2030, yet progress remains off-track with 2.2 billion people lacking these services. Our comprehensive global analysis revealed a striking paradox: 88.5% of those without SMDW resided in regions with abundant rainfall, while only 1.26% used rainwater for potable use. Accordingly, a parametric framework integrating environmental and socioeconomic indicators was developed to quantify rainwater harvesting (RWH)’s potential for advancing SDG 6.1. This framework enabled a stepwise, actionable roadmap centered on two synergistic pathways: extending maximum seasonal storage potential and improving transition ratios, with simulations demonstrating that progressive optimization of RWH could increase SMDW coverage by 5.6–26%, benefiting 0.45–2.08 billion people. Socio-ecological analyses further showed that RWH supported multiple SDGs, including food security, health, gender equality, and climate action. These findings establish a systematic global assessment for unlocking vast untapped rainwater potential, providing important pathways that could fundamentally accelerate the achievement of universal SMDW access.

    Similar content being viewed by others

    Assessing global drinking water potential from electricity-free solar water evaporation device

    Article
    Open access
    08 August 2024

    Global potential for harvesting drinking water from air using solar energy

    Article
    Open access
    27 October 2021

    Decision making for implementing non-traditional water sources: a review of challenges and potential solutions

    Article
    Open access
    10 August 2023

    Data availability

    The primary data used in this study are publicly available from established databases: precipitation data from Climatic Research Unit Timeseries (CRUTS 4.05) at https://crudata.uea.ac.uk/cru/data/hrg/; surface runoff data from Geographic Remote Sensing Ecological Network at https://www.gisrs.cn; water access data from WHO/UNICEF Joint Monitoring Programme at https://washdata.org/; population data from WorldPop at https://www.worldpop.org/; health data from Institute for Health Metrics and Evaluation Global Burden of Disease at http://ghdx.healthdata.org/gbd-results-tool; socioeconomic data from World Bank at https://data.worldbank.org/. The analytical results generated in this study are provided in the Source Data file and Figshare: https://doi.org/10.6084/m9.figshare.30427318 and https://doi.org/10.6084/m9.figshare.30484844. The intermediate processed datasets generated through our analytical framework are available from the corresponding author upon request. Source data are provided with this paper.
    ReferencesGleick, P. H. The human right to water. Water Policy 1, 487–503 (1998).
    Google Scholar 
    Liu, M. J. et al. Spatial assessment of tap-water safety in China. Nat. Sustain. 5, 689–698 (2022).
    Google Scholar 
    Bain, R., Johnston, R. & Slaymaker, T. Drinking water quality and the SDGs. Npj Clean. Water 3, 37 (2020).
    Google Scholar 
    Taka, M. et al. The potential of water security in leveraging Agenda 2030. One Earth 4, 258–268 (2021).
    Google Scholar 
    United Nations Children’s Fund and World Health Organization. Progress on household drinking water, sanitation and hygiene 2000-2022: Special focus on gender. (World Health Organization, 2024).World Health Organization. Progress on household drinking water, sanitation and hygiene 2000-2017: special focus on inequalities. (World Health Organization, 2019).UN-Water. The sustainable development goal 6 global acceleration framework. Geneva: UN-Water (2020).Segovia-Hernández, J. G., Hernández, S., Cossío-Vargas, E. & Sánchez-Ramírez, E. Challenges and opportunities in process intensification to achieve the UN’s 2030 agenda: goals 6, 7, 9, 12 and 13. Chem. Eng. Process. Process Intensif., 109507 (2023).Naylor, K. A. Blueprint for acceleration: sustainable development goal 6 synthesis report on water and sanitation, 2023. (UN, 2023).Hutton, G. & Varughese, M. The costs of meeting the 2030 sustainable development goal targets on drinking water, sanitation, and hygiene. World Bank, Washington, DC (2016).Department of Economic and Social Affairs in the United Nations. Mapping and progress of the UN 2023 Water Conference Water Action Agenda, https://sdgs.un.org/sites/default/files/2024-06/WAA%20Mapping%20and%20Progress%20Report%20May%202024.pdf (2023).Mati, B. et al. Mapping the potential of rainwater harvesting technologies in Africa. A GIS overview on development domains for the continent and ten selected countries. Tech. Man. 6, 126 (2006).
    Google Scholar 
    Mwenge Kahinda, J. & Taigbenu, A. E. Rainwater harvesting in South Africa: Challenges and opportunities. Phys. Chem. Earth 36, 968–976 (2011).
    Google Scholar 
    Wunderlich, S., St George Freeman, S., Galindo, L., Brown, C. & Kumpel, E. Optimizing household water decisions for managing intermittent water supply in Mexico City. Environ. Sci. Technol. 55, 8371–8381 (2021).
    Google Scholar 
    Barron, J. & Salas, J. C. Rainwater harvesting: a lifeline for human well-being. (UNEP/Earthprint, 2009).Alim, M. A. et al. Suitability of roof harvested rainwater for potential potable water production: A scoping review. J. Clean. Prod. 248, 119226 (2020).
    Google Scholar 
    Lebek, K. & Krueger, T. Conventional and makeshift rainwater harvesting in rural South Africa: exploring determinants for rainwater harvesting mode. Int. J. Water Resour. Dev. 39, 113–132 (2023).
    Google Scholar 
    Chaplot, V., Selala, M., Thenga, H. & Jewitt, G. Comparison of the chemical quality of rainwater harvested from roof and surface run-off systems. Water SA 44, 223–231 (2018).
    Google Scholar 
    Ertop, H. et al. The importance of rainwater harvesting and its usage possibilities: Antalya example (Turkey). Water 15, 2194 (2023).
    Google Scholar 
    Dao, D. A. et al. Assessment of rainwater harvesting and maintenance practice for better drinking water quality in rural areas. Aqua Water Infrastruct. Ecosyst. Soc. 70, 202–216 (2021).
    Google Scholar 
    Karim, M. R. Assessment of rainwater harvesting for drinking water supply in Bangladesh. Water Sci. Technol. Water Supply 10, 243–249 (2010).
    Google Scholar 
    Alim, M. A. et al. Feasibility analysis of a small-scale rainwater harvesting system for drinking water production at Werrington, New South Wales, Australia. J. Clean. Prod. 270, 122437 (2020).
    Google Scholar 
    Kahinda, J. -mM., Taigbenu, A. E. & Boroto, J. R. Domestic rainwater harvesting to improve water supply in rural South Africa. Phys. Chem. Earth 32, 1050–1057 (2007).
    Google Scholar 
    Kumar, M. D. Rainwater Harvesting in India: Some Critical Issues for Basin Planning and Research. (2009).Bhatt, S. C. Natural resources, water harvesting and drought in Central India. (Shree Publishers & Distributors, New Delhi, 2020).Qi, Q., Marwa, J., Mwamila, T., Gwenzi, W. & Noubactep, C. Making rainwater harvesting a key solution for water management: the universality of the Kilimanjaro concept. Sustainability 11, 5606 (2019).
    Google Scholar 
    Imteaz, M. A., Adeboye, O. B., Rayburg, S. & Shanableh, A. Rainwater harvesting potential for southwest Nigeria using daily water balance model. Resour. Conserv Recy 62, 51–55 (2012).
    Google Scholar 
    Droogers, P. & van Kempen, C. Mapping the potential for rainwater harvesting under various scenarios. FutureWater Costerweg, Wageningen, the Netherlands (2014).Harris, I., Osborn, T. J., Jones, P. & Lister, D. Version 4 of the CRU TS monthly high-resolution gridded multivariate climate dataset. Sci. Data 7, 109 (2020).
    Google Scholar 
    Rentschler, J., Salhab, M. & Jafino, B. A. Flood exposure and poverty in 188 countries. Nat. Commun. 13, 3527 (2022).
    Google Scholar 
    Fox, S., Agyemang, F., Hawker, L. & Neal, J. Integrating social vulnerability into high-resolution global flood risk mapping. Nat. Commun. 15, 3155 (2024).
    Google Scholar 
    Reed, B. Minimum water quantity needed for domestic uses. New Delhi: World Health Organization Regional Office for South-East Asia (2005).Okoye, C. O., Solyalı, O. & Akıntuğ, B. Optimal sizing of storage tanks in domestic rainwater harvesting systems: A linear programming approach. Resour. Conserv. Recycl. 104, 131–140 (2015).
    Google Scholar 
    Kinkade-Levario, H. Design for water: rainwater harvesting, stormwater catchment, and alternate water reuse. (New Society Publishers, 2007).Campisano, A. & Modica, C. Optimal sizing of storage tanks for domestic rainwater harvesting in Sicily. Resour. Conserv. Recycl. 63, 9–16 (2012).
    Google Scholar 
    Leonard, D. & Gato-Trinidad, S. Understanding the effect of household rainwater tanks on hourly residential water use. J. Sustain. Water Built Environ. 7, 05021004 (2021).
    Google Scholar 
    Cain, N. L. A different path: the global water crisis and rainwater harvesting. Consilience, 147–157 (2014).Ward, S., Memon, F. & Butler, D. Performance of a large building rainwater harvesting system. Water Res 46, 5127–5134 (2012).
    Google Scholar 
    Zabidi, H. A., Goh, H. W., Chang, C. K., Chan, N. W. & Zakaria, N. A. A review of roof and pond rainwater harvesting systems for water security: The design, performance and way forward. Water 12, 3163 (2020).
    Google Scholar 
    Bain, R., Johnston, R., Khan, S., Hancioglu, A. & Slaymaker, T. Monitoring drinking water quality in nationally representative household surveys in low-and middle-income countries: cross-sectional analysis of 27 multiple indicator cluster surveys 2014–2020. Environ. Health Perspect. 129, 097010 (2021).
    Google Scholar 
    Manero, A. et al. Benefits, costs and enabling conditions to achieve ‘water for all’ in rural and remote Australia. Nat. Water 2, 31–40 (2024).
    Google Scholar 
    Nowicki, S., Koehler, J. & Charles, K. J. Including water quality monitoring in rural water services: why safe water requires challenging the quantity versus quality dichotomy. Npj Clean. Water 3, 14 (2020).
    Google Scholar 
    Adeyeye, K. & Bello-Dambatta, A. in Rainwater harvesting for the 21st Century 245–264 (CRC Press).Ye, W. et al. Applicability of photovoltaic panel rainwater harvesting system in improving water-energy-food nexus performance in semi-arid areas. Sci. Total Environ. 896, 164938 (2023).
    Google Scholar 
    Konapala, G., Mishra, A. K., Wada, Y. & Mann, M. E. Climate change will affect global water availability through compounding changes in seasonal precipitation and evaporation. Nat. Commun. 11, 3044 (2020).
    Google Scholar 
    Wolf, J. et al. Burden of disease attributable to unsafe drinking water, sanitation, and hygiene in domestic settings: a global analysis for selected adverse health outcomes. Lancet 401, 2060–2071 (2023).
    Google Scholar 
    Organization, W. H. Burden of disease attributable to unsafe drinking-water, sanitation and hygiene, 2019 update. (World Health Organization, 2023).De Kwaadsteniet, M., Dobrowsky, P., Van Deventer, A., Khan, W. & Cloete, T. Domestic rainwater harvesting: microbial and chemical water quality and point-of-use treatment systems. Water Air Soil Pollut. 224, 1–19 (2013).
    Google Scholar 
    Despins, C., Farahbakhsh, K. & Leidl, C. Assessment of rainwater quality from rainwater harvesting systems in Ontario, Canada. J. Water Supply Res. Technol. AQUA 58, 117–134 (2009).
    Google Scholar 
    Xiong, B. et al. Moringa oleifera f-sand filters for sustainable water purification. Environ. Sci. Technol. Lett. 5, 38–42 (2018).
    Google Scholar 
    Jeon, I., Ryberg, E. C., Alvarez, P. J. & Kim, J.-H. Technology assessment of solar disinfection for drinking water treatment. Nat. Sustain. 5, 801–808 (2022).
    Google Scholar 
    Malina, N., Millner, S. & Ojeda, A. S. “First alert” for drinking water quality: commercial water quality testing kits for community-engaged research. ACS EST Water 4, 1346–1355 (2024).
    Google Scholar 
    Bain, R., Johnston, R., Mitis, F., Chatterley, C. & Slaymaker, T. Establishing sustainable development goal baselines for household drinking water, sanitation and hygiene Services. Water 10 (2018).World Health Organization (WHO) and the United Nations Children’s Fund. JMP Methodology: 2017 update and SDG baselines, https://washdata.org/reports/jmp-2017-methodology (2018).Lord, J. et al. Global potential for harvesting drinking water from air using solar energy. Nature 598, 611–617 (2021).
    Google Scholar 
    FAO. Land and water digital media series, 26. Training course on RWH (CDROM). Planning of water harvesting schemes, Unit 22. Food and Agriculture Organization of the United Nations (2003).Kuzma, S. et al. Aqueduct 4.0: Updated decision-relevant global water risk indicators. (World Resources Institute Washington, DC, USA, 2023).Zhu, Q., Li, Y. & Tang, X. in Rainwater Harvesting for Agriculture and Water Supply (eds Qiang Zhu, John Gould, Yuanhong Li & Chengxiang Ma) 3-42 (Springer Singapore, 2015).Our World in Data team. Water productivity, GDP per cubic meter of freshwater withdrawal, https://ourworldindata.org/grapher/water-productivity (2024).World Bank Group. GDP per capita (current US$), https://data.worldbank.org/indicator/NY.GDP.PCAP.CD (2021).World Bank Group. Rural population, https://data.worldbank.org/indicator/SP.RUR.TOTL (2021).World Bank Group. Population growth (annual %), https://data.worldbank.org/indicator/SP.POP.GROW (2021).Our World in Data team. Prevalence of moderate or severe food insecurity in total population, https://ourworldindata.org/sdgs/zero-hunger (2023).Our World in Data team. Gender inequality index, https://ourworldindata.org/grapher/gender-inequality-index-from-the-human-development-report?time=2020 (2024).Disaster Risk Management Knowledge Centre. DRMKC-INFORM Country Risk Profile, https://drmkc.jrc.ec.europa.eu/inform-index/INFORM-Risk/Country-Risk-Profile (2024).Institute for health metrics and evaluation. Explore results from the 2021 Global Burden of Disease (GBD) study, https://vizhub.healthdata.org/gbd-results/ (2024).Download referencesAcknowledgementsThis research was supported by the Basic Science Center Project of the Natural Science Foundation of China (Grant no. 52388101, C.Z. Hu and F.B. Luan), and China-Sri Lanka Joint Research and Demonstration Center for Water Technology, China-Sri Lanka Joint Center for Education and Research, Chinese Academy of Sciences, China (F.B. Luan).Author informationAuthors and AffiliationsKey Laboratory of Environmental Aquatic Chemistry, State Key Laboratory of Regional Environment and Sustainability, Research Center for Eco-Environmental Sciences, Chinese Academy of Sciences, Beijing, ChinaQingke Yuan, Yanfeng Liu, Yukang Qie, Chengzhi Hu, Ying Meng & Fubo LuanUniversity of Chinese Academy of Sciences, Beijing, ChinaYanfeng Liu, Yukang Qie, Chengzhi Hu & Fubo LuanAuthorsQingke YuanView author publicationsSearch author on:PubMed Google ScholarYanfeng LiuView author publicationsSearch author on:PubMed Google ScholarYukang QieView author publicationsSearch author on:PubMed Google ScholarChengzhi HuView author publicationsSearch author on:PubMed Google ScholarYing MengView author publicationsSearch author on:PubMed Google ScholarFubo LuanView author publicationsSearch author on:PubMed Google ScholarContributionsQ.K. Yuan and F.B. Luan conceived and designed the study, developed the parametric framework linking rainwater harvesting potential to SDG 6.1 indicators, and prepared the initial manuscript. Y. Meng refined the parameter selection and normalization processes, constructed simulation scenarios, and assisted with data verification and quality control. Y.F. Liu and Y.K. Qie conducted data acquisition, analysis, and modeling, integrating multi-source global datasets, and created visualizations to illustrate key findings. Q.K. Yuan completed the manuscript writing. C.Z. Hu developed the transition ratio pathway within the roadmap framework, focusing on water quality requirements and quality management strategies. All authors participated in reviewing, editing, and refining the manuscript. F.B. Luan supervised the overall research, providing guidance and oversight throughout the study.Corresponding authorsCorrespondence to
    Ying Meng or Fubo Luan.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Peer review

    Peer review information
    Nature Communications thanks Robert Bain and Ataur Rahman for their contribution to the peer review of this work. A peer review file is available.

    Additional informationPublisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary informationSupplementary InformationTransparent Peer Review fileSource dataSource dataRights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleYuan, Q., Liu, Y., Qie, Y. et al. Unlocking global rainwater harvesting potential for safe drinking water access.
    Nat Commun (2025). https://doi.org/10.1038/s41467-025-66429-wDownload citationReceived: 07 February 2025Accepted: 05 November 2025Published: 12 December 2025DOI: https://doi.org/10.1038/s41467-025-66429-wShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More