More stories

  • in

    Accelerated land surface greening caused by earlier permafrost thawing

    AbstractPersistent and above-average warming has advanced the start of spring permafrost thawing, intensifying climate warming through carbon feedbacks. However, the extent to which variations in spring permafrost thawing contribute to greening trends in permafrost-affected areas (i.e., increases in vegetation greenness) over time remains unclear, limiting our understanding of the ecological consequences of permafrost degradation. Analyzing 40-year freeze/thaw data and multiple satellite-derived greenness indicators, we identify widespread increases in the sensitivity of spring greenness to spring permafrost thawing based on moving-window analyses, indicating that advances in spring permafrost thawing have played a progressively stronger role in promoting spring greening, particularly in boreal forests and tundra regions underlain by continuous permafrost. In addition to the region-specific climate and permafrost conditions, we uncover biogeophysical pathways accounting for the increase in sensitivity of spring greenness to spring permafrost thawing, including reduced albedo, earlier vegetation phenology, and enhanced soil moisture infiltration. Notably, state-of-the-art Dynamic Global Vegetation Models consistently underestimate both the magnitude and variability of sensitivity of spring greenness to spring permafrost thawing. These findings highlight the temporal changes in vegetation responses to freeze-thaw dynamics, necessitating improved model projections concurrent with climate change.

    Similar content being viewed by others

    Tundra vegetation change and impacts on permafrost

    Article

    11 January 2022

    Vegetation greening amplifies shallow soil temperature warming on the Tibetan Plateau

    Article
    Open access
    04 June 2024

    Warming-independent shortened snow cover duration enhances vegetation greening across northern permafrost region

    Article
    Open access
    01 April 2025

    Data availability

    All data used in this study are freely available from public repositories. Freeze-thaw state data are provided by FT-ESDR (https://nsidc.org/data/nsidc-0477/versions/5). Satellite-based greenness indicators include kNDVI (https://www.environment.snu.ac.kr/data/longterm-vi), LCSPP v3.2 (https://zenodo.org/records/14568024), BESS v2.0 GPP (https://www.environment.snu.ac.kr/bessv2), GOSIF, CSIF, and VODCA v2 (https://researchdata.tuwien.at/records/t74ty-tcx62). Meteorological data are available from TerraClimate (https://www.climatologylab.org/). Global monthly CO2 records are available from NOAA. ERA5-Land datasets are available from Google Earth Engine and Copernicus Climate Data Store (https://cds.climate.copernicus.eu/). In situ surface temperature records are available from GTN-P (https://gtnp.arcticportal.org/). Active layer thickness data are provided by ESA. Soil organic carbon and nitrogen data are available from SoilGrids (https://www.isric.org/explore/soilgrids). Maximum root depth is available from EartH2Observe (http://www.earth2observe.eu/). Global canopy height data are available from ETH Global Canopy Height 2020 (https://code.earthengine.google.com/126c172d63e7ce780596c8d26f06d384). Land cover datasets include TEOW biomes (https://www.worldwildlife.org/publications/terrestrial-ecoregions-of-the-world), MODIS MCD12C1 IGBP (https://lpdaac.usgs.gov/), MOSEV dNBR (https://zenodo.org/records/4265209), and permafrost type data from NSIDC. Source data are provided with this paper.
    Code availability

    All data analyses were performed using R v.4.3.1. The codes used in this study are available at Zenodo [https://doi.org/10.5281/zenodo.17543943]95.
    ReferencesHeijmans, M. M. P. D. Tundra vegetation change and impacts on permafrost. Nat. Rev. Earth Environ. 3, 68–84 (2022).Myneni, R. B., Keeling, C. D., Tucker, C. J., Asrar, G. & Nemani, R. R. Increased plant growth in the northern high latitudes from 1981 to 1991. Nature 386, 698–702 (1997).
    Google Scholar 
    Zhu, Z. et al. Greening of the Earth and its drivers. Nat. Clim. Change 6, 791–795 (2016).
    Google Scholar 
    Pearson, R. G. et al. Shifts in Arctic vegetation and associated feedbacks under climate change. Nat. Clim. Change 3, 673–677 (2013).
    Google Scholar 
    Piao, S. et al. Characteristics, drivers and feedbacks of global greening. Nat. Rev. Earth Environ. 1, 14–27 (2020).
    Google Scholar 
    Xu, X., Riley, W. J., Koven, C. D., Jia, G. & Zhang, X. Earlier leaf-out warms air in the north. Nat. Clim. Change 10, 370–375 (2020).
    Google Scholar 
    Schuur, E. A. G. et al. Climate change and the permafrost carbon feedback. Nature 520, 171–179 (2015).
    Google Scholar 
    Walvoord, M. A. & Kurylyk, B. L. Hydrologic impacts of thawing permafrost—a review. Vadose Zone J. 15, 1–20 (2016).
    Google Scholar 
    Webb, E. E. et al. Permafrost thaw drives surface water decline across lake-rich regions of the Arctic. Nat. Clim. Change 12, 841–846 (2022).
    Google Scholar 
    Biskaborn, B. K. et al. Permafrost is warming at a global scale. Nat. Commun. 10, 264 (2019).
    Google Scholar 
    Smith, S. L., O’Neill, H. B., Isaksen, K., Noetzli, J. & Romanovsky, V. E. The changing thermal state of permafrost. Nat. Rev. Earth Environ. 3, 10–23 (2022).
    Google Scholar 
    Natali, S. M., Schuur, E. A. G. & Rubin, R. L. Increased plant productivity in Alaskan tundra as a result of experimental warming of soil and permafrost. J. Ecol. 100, 488–498 (2012).
    Google Scholar 
    Schuur, E. A. G. & Mack, M. C. Ecological response to permafrost Thaw and consequences for local and global ecosystem services. Annu. Rev. Ecol. Evol. Syst. 49, 279–301 (2018).
    Google Scholar 
    Luo, S., Wang, J., Pomeroy, J. W. & Lyu, S. Freeze–Thaw changes of seasonally frozen ground on the Tibetan Plateau from 1960 to 2014. J. Clim. 33, 9427–9446 (2020).
    Google Scholar 
    Li, X., Jin, R., Pan, X., Zhang, T. & Guo, J. Changes in the near-surface soil freeze–thaw cycle on the Qinghai-Tibetan Plateau. Int. J. Appl. Earth Observ. Geoinf. 17, 33–42 (2012).
    Google Scholar 
    Wang, J. & Liu, D. Vegetation green-up date is more sensitive to permafrost degradation than climate change in spring across the northern permafrost region. Glob. Change Biol. 28, 1569–1582 (2022).
    Google Scholar 
    Jin, X.-Y. et al. Impacts of climate-induced permafrost degradation on vegetation: a review. Adv. Clim. Change Res. 12, 29–47 (2021).
    Google Scholar 
    Wrona, F. J. et al. Transitions in Arctic ecosystems: Ecological implications of a changing hydrological regime. JGR Biogeosci. 121, 650–674 (2016).
    Google Scholar 
    Yang, M., Nelson, F. E., Shiklomanov, N. I., Guo, D. & Wan, G. Permafrost degradation and its environmental effects on the Tibetan Plateau: a review of recent research. Earth-Sci. Rev. 103, 31–44 (2010).
    Google Scholar 
    Buermann, W. et al. Widespread seasonal compensation effects of spring warming on northern plant productivity. Nature 562, 110–114 (2018).
    Google Scholar 
    Piao, S. et al. Evidence for a weakening relationship between interannual temperature variability and northern vegetation activity. Nat. Commun. 5, 5018 (2014).
    Google Scholar 
    Jorgenson, M. T., Shur, Y. L. & Pullman, E. R. Abrupt increase in permafrost degradation in Arctic Alaska. Geophys. Res. Lett. 33, 2005GL024960 (2006).
    Google Scholar 
    Berner, L. T. et al. Summer warming explains widespread but not uniform greening in the Arctic tundra biome. Nat. Commun. 11, 4621 (2020).
    Google Scholar 
    Grünberg, I., Wilcox, E. J., Zwieback, S., Marsh, P. & Boike, J. Linking tundra vegetation, snow, soil temperature, and permafrost. Biogeosciences 17, 4261–4279 (2020).
    Google Scholar 
    Miller, P. A. & Smith, B. Modelling tundra vegetation response to recent Arctic warming. AMBIO 41, 281–291 (2012).
    Google Scholar 
    Tang, S. et al. Resonance between projected Tibetan Plateau surface darkening and Arctic climate change. Sci. Bull. 69, 367–374 (2024).
    Google Scholar 
    Wang, J. & Liu, D. Deciphering the biophysical impact of permafrost greening on summer surface offset. Earth’s. Future 12, e2023EF004077 (2024).
    Google Scholar 
    Pistone, K., Eisenman, I. & Ramanathan, V. Observational determination of albedo decrease caused by vanishing Arctic sea ice. Proc. Natl. Acad. Sci. USA 111, 3322–3326 (2014).
    Google Scholar 
    Dragoni, D. et al. Evidence of increased net ecosystem productivity associated with a longer vegetated season in a deciduous forest in south-central Indiana, USA. Glob. Change Biol. 17, 886–897 (2011).
    Google Scholar 
    Richardson, A. D. et al. Influence of spring phenology on seasonal and annual carbon balance in two contrasting New England forests. Tree Physiol. 29, 321–331 (2009).
    Google Scholar 
    Li, J. et al. Weakening warming on spring freeze–thaw cycle caused greening Earth’s third pole. Proc. Natl. Acad. Sci. USA 121, e2319581121 (2024).
    Google Scholar 
    Bui, M. T., Lu, J. & Nie, L. A review of hydrological models applied in the permafrost-dominated Arctic region. Geosciences 10, 401 (2020).
    Google Scholar 
    Shu, S., Jain, A. K., Koven, C. D. & Mishra, U. Estimation of permafrost SOC stock and turnover time using a land surface model with vertical heterogeneity of permafrost soils. Glob. Biogeochem. Cycles 34, e2020GB006585 (2020).
    Google Scholar 
    Wang, T. et al. Permafrost thawing puts the frozen carbon at risk over the Tibetan Plateau. Sci. Adv. 6, eaaz3513 (2020).
    Google Scholar 
    Hjort, J. et al. Impacts of permafrost degradation on infrastructure. Nat. Rev. Earth Environ. 3, 24–38 (2022).
    Google Scholar 
    Loranty, M. M. et al. Reviews and syntheses: changing ecosystem influences on soil thermal regimes in northern high-latitude permafrost regions. Biogeosciences 15, 5287–5313 (2018).
    Google Scholar 
    Turetsky, M. R. et al. Permafrost collapse is accelerating carbon release. Nature 569, 32–34 (2019).
    Google Scholar 
    Strauss, J. et al. Deep Yedoma permafrost: a synthesis of depositional characteristics and carbon vulnerability. Earth-Sci. Rev. 172, 75–86 (2017).
    Google Scholar 
    Bintanja, R. & Andry, O. Towards a rain-dominated Arctic. Nat. Clim. Change 7, 263–267 (2017).
    Google Scholar 
    Liljedahl, A. K. et al. Pan-Arctic ice-wedge degradation in warming permafrost and its influence on tundra hydrology. Nat. Geosci. 9, 312–318 (2016).
    Google Scholar 
    Neumann, R. B. et al. Warming effects of spring rainfall increase methane emissions from thawing permafrost. Geophys. Res. Lett. 46, 1393–1401 (2019).
    Google Scholar 
    Pilyugina, P. et al. A physics-informed machine learning framework for permafrost stability assessment. IEEE Access 13, 96423–96433 (2025).
    Google Scholar 
    Kim, Y., Kimball, J. S., Glassy, J. & Du, J. An extended global Earth system data record on daily landscape freeze–thaw status determined from satellite passive microwave remote sensing. Earth Syst. Sci. Data 9, 133–147 (2017).
    Google Scholar 
    Muñoz-Sabater, J. et al. ERA5-Land: a state-of-the-art global reanalysis dataset for land applications. Earth Syst. Sci. Data 13, 4349–4383 (2021).
    Google Scholar 
    Chen, X. et al. Different responses of surface freeze and thaw phenology changes to warming among Arctic permafrost types. Remote Sens. Environ. 272, 112956 (2022).
    Google Scholar 
    Kim, Y., Kimball, J. S., Zhang, K. & McDonald, K. C. Satellite detection of increasing Northern Hemisphere non-frozen seasons from 1979 to 2008: implications for regional vegetation growth. Remote Sens. Environ. 121, 472–487 (2012).
    Google Scholar 
    Li, J., Wu, C., Peñuelas, J., Ran, Y. & Zhang, Y. The start of frozen dates over northern permafrost regions with the changing climate. Glob. Change Biol. 29, 4556–4568 (2023).
    Google Scholar 
    Camps-Valls, G. et al. A unified vegetation index for quantifying the terrestrial biosphere. Sci. Adv. 7, eabc7447 (2021).
    Google Scholar 
    Chen, J. M. et al. Vegetation structural change since 1981 significantly enhanced the terrestrial carbon sink. Nat. Commun. 10, 4259 (2019).
    Google Scholar 
    Porcar-Castell, A. et al. Chlorophyll a fluorescence illuminates a path connecting plant molecular biology to Earth-system science. Nat. Plants 7, 998–1009 (2021).
    Google Scholar 
    Hu, Z. et al. Decoupling of greenness and gross primary productivity as aridity decreases. Remote Sens. Environ. 279, 113120 (2022).
    Google Scholar 
    Ding, Z., Peng, J., Qiu, S. & Zhao, Y. Nearly half of global vegetated area experienced inconsistent vegetation growth in terms of greenness, cover, and productivity. Earth’s. Future 8, e2020EF001618 (2020).
    Google Scholar 
    Jeong, S. et al. Persistent global greening over the last four decades using novel long-term vegetation index data with enhanced temporal consistency. Remote Sens. Environ. 311, 114282 (2024).
    Google Scholar 
    Fang, J. et al. A long-term reconstruction of a global photosynthesis proxy over 1982–2023. Sci. Data 12, 372 (2025).
    Google Scholar 
    Lian, X. et al. Diminishing carryover benefits of earlier spring vegetation growth. Nat. Ecol. Evol. 8, 218–228 (2024).
    Google Scholar 
    Li, B. et al. BESSv2.0: A satellite-based and coupled-process model for quantifying long-term global land–atmosphere fluxes. Remote Sens. Environ. 295, 113696 (2023).
    Google Scholar 
    Li, X. & Xiao, J. A global, 0.05-degree product of solar-induced chlorophyll fluorescence derived from OCO-2, MODIS, and reanalysis data. Remote Sens. 11, 517 (2019).
    Google Scholar 
    Zhang, Y., Joiner, J., Alemohammad, S. H., Zhou, S. & Gentine, P. A global spatially contiguous solar-induced fluorescence (CSIF) dataset using neural networks. Biogeosciences 15, 5779–5800 (2018).
    Google Scholar 
    Zotta, R.-M. et al. VODCA v2: multi-sensor, multi-frequency vegetation optical depth data for long-term canopy dynamics and biomass monitoring. Earth Syst. Sci. Data 16, 4573–4617 (2024).
    Google Scholar 
    Friedlingstein, P. et al. Global Carbon Budget 2023. Earth Syst. Sci. Data 15, 5301–5369 (2023).
    Google Scholar 
    Sitch, S. et al. Recent trends and drivers of regional sources and sinks of carbon dioxide. Biogeosciences 12, 653–679 (2015).
    Google Scholar 
    Abatzoglou, J. T., Dobrowski, S. Z., Parks, S. A. & Hegewisch, K. C. TerraClimate, a high-resolution global dataset of monthly climate and climatic water balance from 1958–2015. Sci. Data 5, 170191 (2018).
    Google Scholar 
    Li, W. et al. Widespread increasing vegetation sensitivity to soil moisture. Nat. Commun. 13, 3959 (2022).
    Google Scholar 
    Pinzon, J. & Tucker, C. A non-stationary 1981–2012 AVHRR NDVI3g time series. Remote Sens. 6, 6929–6960 (2014).
    Google Scholar 
    Shen, M. et al. Increasing altitudinal gradient of spring vegetation phenology during the last decade on the Qinghai–Tibetan Plateau. Agric. For. Meteorol. 189–190, 71–80 (2014).
    Google Scholar 
    Chen, J. et al. A simple method for reconstructing a high-quality NDVI time-series data set based on the Savitzky–Golay filter. Remote Sens. Environ. 91, 332–344 (2004).
    Google Scholar 
    Wang, J., Liu, D., Ciais, P. & Peñuelas, J. Decreasing rainfall frequency contributes to earlier leaf onset in northern ecosystems. Nat. Clim. Change 12, 386–392 (2022).
    Google Scholar 
    Wu, C. et al. Increased drought effects on the phenology of autumn leaf senescence. Nat. Clim. Change https://doi.org/10.1038/s41558-022-01464-9 (2022).
    Google Scholar 
    Wang, X. & Wu, C. Estimating the peak of growing season (POS) of China’s terrestrial ecosystems. Agric. For. Meteorol. 278, 107639 (2019).
    Google Scholar 
    Fan, Y., Miguez-Macho, G., Jobbágy, E. G., Jackson, R. B. & Otero-Casal, C. Hydrologic regulation of plant rooting depth. Proc. Natl. Acad. Sci. USA 114, 10572–10577 (2017).
    Google Scholar 
    Lang, N., Jetz, W., Schindler, K. & Wegner, J. D. A high-resolution canopy height model of the Earth. Nat. Ecol. Evol. 7, 1778–1789 (2023).
    Google Scholar 
    Obu, J. et al. Northern Hemisphere permafrost map based on TTOP modelling for 2000–2016 at 1 km2 scale. Earth-Sci. Rev. 193, 299–316 (2019).
    Google Scholar 
    Poggio, L. et al. SoilGrids 2.0: producing soil information for the globe with quantified spatial uncertainty. Soil 7, 217–240 (2021).
    Google Scholar 
    Alonso-González, E. & Fernández-García, V. MOSEV: a global burn severity database from MODIS (2000–2020). Earth Syst. Sci. Data 13, 1925–1938 (2021).
    Google Scholar 
    Brown, J., Ferrians, O., Heginbottom, J. A. & Melnikov, E. Circum-Arctic map of permafrost and ground-ice conditions. Version 2. 10.7265/skbg-kf16 (2002).Chen, L. et al. Leaf senescence exhibits stronger climatic responses during warm than during cold autumns. Nat. Clim. Change 10, 777–780 (2020).
    Google Scholar 
    He, L. et al. Lagged precipitation effects on plant production across terrestrial biomes. Nat. Ecol. Evol. https://doi.org/10.1038/s41559-025-02806-4 (2025).
    Google Scholar 
    Wang, S. et al. Recent global decline of CO 2 fertilization effects on vegetation photosynthesis. Science 370, 1295–1300 (2020).
    Google Scholar 
    Fu, Y. H. et al. Declining global warming effects on the phenology of spring leaf unfolding. Nature 526, 104–107 (2015).
    Google Scholar 
    Yang, G., Crowther, T. W., Lauber, T., Zohner, C. M. & Smith, G. R. A globally consistent negative effect of edge on aboveground forest biomass. Nat. Ecol. Evol. https://doi.org/10.1038/s41559-025-02840-2 (2025).
    Google Scholar 
    Wolkovich, E. M. et al. A simple explanation for declining temperature sensitivity with warming. Glob. Change Biol. 27, 4947–4949 (2021).
    Google Scholar 
    Breiman, L. Random forests. Mach. Learn. 45, 5–32 (2001).
    Google Scholar 
    Strobl, C., Boulesteix, A.-L., Kneib, T., Augustin, T. & Zeileis, A. Conditional variable importance for random forests. BMC Bioinform. 9, 307 (2008).
    Google Scholar 
    Wang, X. et al. Enhanced habitat loss of the Himalayan endemic flora driven by warming-forced upslope tree expansion. Nat. Ecol. Evol. 6, 890–899 (2022).
    Google Scholar 
    Berdugo, M., Gaitán, J. J., Delgado-Baquerizo, M., Crowther, T. W. & Dakos, V. Prevalence and drivers of abrupt vegetation shifts in global drylands. Proc. Natl. Acad. Sci. USA 119, e2123393119 (2022).
    Google Scholar 
    Wright, M. N. & Ziegler, A. ranger: a fast implementation of random forests for high dimensional data in C++ and R. J. Stat. Soft. 77, 1–17 (2017).Lundberg, S. M. et al. From local explanations to global understanding with explainable AI for trees. Nat. Mach. Intell. 2, 56–67 (2020).Keenan, T. F. et al. Net carbon uptake has increased through warming-induced changes in temperate forest phenology. Nat. Clim. Change 4, 598–604 (2014).
    Google Scholar 
    Sugihara, G. et al. Detecting causality in complex ecosystems. Science 338, 496–500 (2012).
    Google Scholar 
    Shen, M. et al. Can changes in autumn phenology facilitate earlier green-up date of northern vegetation?. Agric. For. Meteorol. 291, 108077 (2020).
    Google Scholar 
    Wu, C. et al. Contrasting responses of autumn-leaf senescence to daytime and night-time warming. Nat. Clim. Change 8, 1092–1096 (2018).
    Google Scholar 
    Bagozzi, R. P. & Yi, Y. Specification, evaluation, and interpretation of structural equation models. J. Acad. Mark. Sci. 40, 8–34 (2012).
    Google Scholar 
    Gao, S. et al. An earlier start of the thermal growing season enhances tree growth in cold humid areas but not in dry areas. Nat. Ecol. Evol. 6, 397–404 (2022).
    Google Scholar 
    Rosseel, Y. lavaan: an R package for structural equation modeling. J. Stat. Soft. 48, 1–36 (2012).Hua, H., Wang, J. & Wu, C. Accelerated land surface greening caused by earlier permafrost thawing. Zenodo 10.5281/ZENODO.17543942 (2025).Download referencesAcknowledgementsThis work was funded by the National Natural Science Foundation of China (42125101, W2412014). J.W. was funded by the National Natural Science Foundation of China (42571037) and “Kezhen and Bingwei” Young Scientist Program of IGSNRR. C.M.Z. was funded by SNF Ambizione grant PZ00P3_193646. J.P. was funded by the TED2021-132627B-I00 grant funded by the Spanish MCIN, AEI/10.13039/501100011033 and by the European Union NextGenerationEU/PRTR, the Fundación Ramón Areces project CIVP20A6621 and the Catalan government grants SGR221-1333 and AGAUR2023 CLIMA 00118. We also appreciate the funding from the Science and Technology Program of Guangdong (No. 2024B1212070012).Author informationAuthor notesThese authors contributed equally: Hao Hua, Jian Wang.Authors and AffiliationsThe Key Laboratory of Land Surface Pattern and Simulation, Institute of Geographical Sciences and Natural Resources Research, Chinese Academy of Sciences, Beijing, ChinaHao Hua, Jian Wang & Chaoyang WuUniversity of the Chinese Academy of Sciences, Beijing, ChinaHao Hua, Jian Wang & Chaoyang WuDepartment of Environmental Systems Science, Institute of Integrative Biology, ETH, Zurich, Zurich, SwitzerlandConstantin M. ZohnerCSIC, Global Ecology Unit CREAF-CSIC-UAB, Barcelona, Catalonia, SpainJosep PeñuelasCREAF, Barcelona, Catalonia, SpainJosep PeñuelasNorthwest Institute of Eco-Environment and Resources, Chinese Academy of Sciences, Lanzhou, ChinaYouhua RanAuthorsHao HuaView author publicationsSearch author on:PubMed Google ScholarJian WangView author publicationsSearch author on:PubMed Google ScholarConstantin M. ZohnerView author publicationsSearch author on:PubMed Google ScholarJosep PeñuelasView author publicationsSearch author on:PubMed Google ScholarYouhua RanView author publicationsSearch author on:PubMed Google ScholarChaoyang WuView author publicationsSearch author on:PubMed Google ScholarContributionsC.W. designed the research. H.H. and J.W. wrote the first draft of the manuscript and performed the analyses. C.M.Z. and J.P. discussed the research design, interpretation, and manuscript revision. Y.R. provided methodological support. All authors assessed the analyses and contributed to improving the manuscript.Corresponding authorCorrespondence to
    Chaoyang Wu.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Peer review

    Peer review information
    Nature Communications thanks Andreas Dietz, who co-reviewed with Martina Wenzl; Steven Cumming, and Tingting Xu 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 SummaryPeer 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 articleHua, H., Wang, J., Zohner, C.M. et al. Accelerated land surface greening caused by earlier permafrost thawing.
    Nat Commun (2025). https://doi.org/10.1038/s41467-025-67644-1Download citationReceived: 13 January 2025Accepted: 04 December 2025Published: 16 December 2025DOI: https://doi.org/10.1038/s41467-025-67644-1Share 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

    Climate and sedimentary structure drive deep labile carbon accumulation in alpine wetlands

    AbstractWetlands serve as vital carbon sinks; however, compared with the surface carbon pool, the relative stock and stability of organic carbon (OC) buried in deep sediment layers remain uncertain, particularly in alpine regions. Based on 42 sediment across seven wetlands on the Qinghai-Xizang Plateau, this study disentangled the OC accumulation process and its drivers since the Holocene. Our results highlighted that deep sediments ( >1 m) stored ~70% of the total OC in alpine wetlands, much of which was labile. Historical warming facilitated the accumulation of such labile OC. Moreover, an intercalated sedimentary structure, formed by silt and fine-grained clay both below and above the OC-rich layer, prevented them from rapid decomposition. Given that elevated groundwater temperatures and intensified hydrological processes in alpine regions may stimulate the decomposition of these massive labile OC pools, releasing carbon into surface water or the atmosphere, future climate change assessments should take this long-overlooked carbon pool into account.

    Similar content being viewed by others

    Warming drives dissolved organic carbon export from pristine alpine soils

    Article
    Open access
    25 April 2024

    Substantial increase of organic carbon storage in Chinese lakes

    Article
    Open access
    14 September 2024

    The land–ocean Arctic carbon cycle

    Article

    06 February 2025

    Data availability

    The data supporting the findings of this study were deposited on figshare (https://doi.org/10.6084/m9.figshare.30655583).
    Code availability

    The R packages utilized in this study are publicly available, and no custom code was used.
    ReferencesWhiting, G. J. & Chanton, J. P. Greenhouse carbon balance of wetlands: methane emission versus carbon sequestration. Tellus Ser. B-Chem. Phys. Meteorol. 53, 521–528 (2001).
    Google Scholar 
    Stewart, A. J. et al. Revealing the hidden carbon in forested wetland soils. Nat. Commun. 15, 726 (2024).
    Google Scholar 
    Jackson, R. B. et al. The ecology of soil carbon: pools, vulnerabilities, and biotic and abiotic controls. Annu. Rev. Ecol. Evol. Syst. 48, 419–445 (2017).
    Google Scholar 
    Li, Y. et al. Factors controlling peat soil thickness and carbon storage in temperate peatlands based on UAV high-resolution remote sensing. Geoderma 449, 117009 (2024).
    Google Scholar 
    Hinson, A. L. et al. The spatial distribution of soil organic carbon in tidal wetland soils of the continental United States. Glob. Change Biol. 23, 5468–5480 (2017).
    Google Scholar 
    Ding, J. Z. et al. The permafrost carbon inventory on the Tibetan Plateau: a new evaluation using deep sediment cores. Glob. Change Biol. 22, 2688–2701 (2016).
    Google Scholar 
    Post, W. M., Emanuel, W. R., Zinke, P. J. & Stangenberger, A. G. Soil carbon pools and world life zones. Nature 298, 156–159 (1982).
    Google Scholar 
    Gitay, H. et al. in Contribution of Working Group II to the IPCC Third Assessment Report: Impacts, Adaptation and Vulnerability (eds J. J. McCarthy et al.) 235-342 (Cambridge University Press, 2001).Vasilevich, R., Lodygin, E., Beznosikov, V. & Abakumov, E. Molecular composition of raw peat and humic substances from permafrost peat soils of European Northeast Russia as climate change markers. Sci. Total Environ. 615, 1229–1238 (2018).
    Google Scholar 
    Routh, J. et al. Multi-proxy study of soil organic matter dynamics in permafrost peat deposits reveal vulnerability to climate change in the European Russian Arctic. Chem. Geol. 368, 104–117 (2014).
    Google Scholar 
    Kalisz, B., Urbanowicz, P., Smólczyński, S. & Orzechowski, M. Impact of siltation on the stability of organic matter in drained peatlands. Ecol. Indic. 130, 108149 (2021).
    Google Scholar 
    Negassa, W., Acksel, A., Eckhardt, K.-U., Regier, T. & Leinweber, P. Soil organic matter characteristics in drained and rewetted peatlands of northern Germany: Chemical and spectroscopic analyses. Geoderma 353, 468–481 (2019).
    Google Scholar 
    Davies, M. A., Blewett, J., Naafs, B. D. A. & Finkelstein, S. A. Ecohydrological controls on apparent rates of peat carbon accumulation in a boreal bog record from the Hudson Bay Lowlands, northern Ontario, Canada. Quat. Res. 104, 14–27 (2021).
    Google Scholar 
    Gao, C. et al. High intensity fire accelerates accumulation of a stable carbon pool in permafrost peatlands under climate warming. Catena 227, 107108 (2023).
    Google Scholar 
    Zhang, T. et al. Warming-driven erosion and sediment transport in cold regions. Nat. Rev. Earth Environ. 3, 832–851 (2022).
    Google Scholar 
    García Lino, M. C. et al. Carbon dynamics in high-Andean tropical cushion peatlands: A review of geographic patterns and potential drivers. Ecol. Monogr. 94, e1614 (2024).
    Google Scholar 
    Heger, A., Becker, J. N., Vásconez, L. K. & Eschenbach, A. Drivers for soil organic carbon stabilization in Elbe River floodplains. J. Plant Nutr. Soil Sci. 187, 346–355 (2024).
    Google Scholar 
    Deiss, L. et al. Soil carbon fractions from an alluvial soil texture gradient in North Carolina. Soil Sci. Soc. Am. J. 81, 1096–1106 (2017).
    Google Scholar 
    Chen, H. et al. Carbon and nitrogen cycling on the Qinghai–Tibetan Plateau. Nat. Rev. Earth Environ. 3, 701–716 (2022).
    Google Scholar 
    Jones, M. C. & Yu, Z. Rapid deglacial and early Holocene expansion of peatlands in Alaska. Proc. Natl. Acad. Sci. USA 107, 7347–7352 (2010).
    Google Scholar 
    Loisel, J. et al. Insights and issues with estimating northern peatland carbon stocks and fluxes since the Last Glacial Maximum. Earth-Sci. Rev. 165, 59–80 (2017).
    Google Scholar 
    Juselius-Rajamäki, T., Vaeliranta, M. & Korhola, A. The ongoing lateral expansion of peatlands in Finland. Glob. Change Biol. 29, 7173–7191 (2023).
    Google Scholar 
    Gao, J., Yao, T. D., Masson-Delmotte, V., Steen-Larsen, H. C. & Wang, W. C. Collapsing glaciers threaten Asia’s water supplies. Nature 565, 19–21 (2019).
    Google Scholar 
    Yao, T. et al. The imbalance of the Asian water tower. Nat. Rev. Earth Environ. 3, 618–632 (2022).
    Google Scholar 
    Feldman, A. F. et al. Plant responses to changing rainfall frequency and intensity. Nat. Rev. Earth Environ. 5, 276–294 (2024).
    Google Scholar 
    Shi, P. J. et al. Earthquakes have accelerated the carbon dioxide emission rate of soils on the Qinghai-Tibet Plateau. Glob. Change Biol. 31, 10 (2025).
    Google Scholar 
    Jobbágy, E. G. & Jackson, R. B. The vertical distribution of soil organic carbon and its relation to climate and vegetation. Ecol. Appl. 10, 423–436 (2000).
    Google Scholar 
    Chen, H. et al. The carbon stock of alpine peatlands on the Qinghai–Tibetan Plateau during the Holocene and their future fate. Quat. Sci. Rev. 95, 151–158 (2014).
    Google Scholar 
    Krauss, K. W. et al. The role of the upper Tidal Estuary in Wetland blue carbon storage and flux. Glob. Biogeochem. Cycles 32, 817–839 (2018).
    Google Scholar 
    Wang, Q. F. et al. The vertical distribution of soil organic carbon and nitrogen in a permafrost-affected wetland on the Qinghai-Tibet Plateau: implications for Holocene development and environmental change. Permafr. Periglac. Process 33, 286–297 (2022).
    Google Scholar 
    Batjes, N. H. Total carbon and nitrogen in the soils of the world. Eur. J. Soil Sci. 65, 10–21 (2014).
    Google Scholar 
    Şenkul, Ç, Gürboğa, Ş, Doğan, M. & Doğan, T. High-resolution geochemical (μXRF) and palynological analyses for climatic and environmental changes in lake sediments from Sultansazlığı Marsh (Central Anatolia) during the last 14.5 kyr. Quat. Int. 613, 24–38 (2022).
    Google Scholar 
    Sun, Z. et al. Changes in atmospheric circulation and glacier melting since the last deglaciation revealed by a lacustrine dD record at Ngamring Co, the upper-middle Yarlung Tsangpo watershed. Paleogeogr. Paleoclimatol. Paleoecol. 598, 11 (2022).
    Google Scholar 
    Gu, Z., Liu, J., Yuan, B. & Liu, D. Monsoon variations of the Qinghai-Xizang Plateau during the last 12,000 years——geochemical evidence from the sediments in the Siling lake. Chin. Sci. Bull. 38, 577–581 (1993).
    Google Scholar 
    Wang, Y., Xia, A. & Xue, K. Cold and humid climates enrich soil carbon stock in the Third Pole grasslands. Innovation 5, 100545 (2024).
    Google Scholar 
    Liu, L. N. et al. Spatial and temporal variations of vegetation cover on the central and eastern Tibetan Plateau since the Last Glacial Period. Glob. Planet. Change 240, 14 (2024).
    Google Scholar 
    Bond, G. et al. Persistent solar influence on North Atlantic climate during the Holocene. Science 294, 2130–2136 (2001).
    Google Scholar 
    Peltre, C., Bruun, S., Du, C. W., Thomsen, I. K. & Jensen, L. S. Assessing soil constituents and labile soil organic carbon by mid-infrared photoacoustic spectroscopy. Soil Biol. Biochem. 77, 41–50 (2014).
    Google Scholar 
    Harris, L. I. et al. Permafrost thaw causes large carbon loss in boreal peatlands while changes to peat quality are limited. Glob. Chang Biol. 29, 5720–5735 (2023).
    Google Scholar 
    Martínez Cortizas, A. et al. 9000 years of changes in peat organic matter composition in Store Mosse (Sweden) traced using FTIR-ATR. Boreas 50, 1161–1178 (2021).
    Google Scholar 
    Oades, J. M. The retention of organic-matter in soils. Biogeochemistry 5, 35–70 (1988).
    Google Scholar 
    Jacobs, P. M. & Mason, J. A. Impact of Holocene dust aggradation on A horizon characteristics and carbon storage in loess-derived Mollisols of the Great Plains, USA. Geoderma 125, 95–106 (2005).
    Google Scholar 
    Yamamoto, K. et al. Tropical peat debris storage in the tidal flat in northern part of the Bengkalis island, Indonesia. MATEC Web Conf. 276, 06002 (2019).
    Google Scholar 
    Yang, S. et al. Burial of organic carbon in Holocene sediments of the Zhujiang (Pearl River) and Changjiang (Yangtze River) estuaries. Mar. Chem. 123, 1–10 (2011).
    Google Scholar 
    Zhou, G. Y. et al. Climate and litter C/N ratio constrain soil organic carbon accumulation. Natl. Sci. Rev. 6, 746–757 (2019).
    Google Scholar 
    Li, Q.-W. et al. Precipitation patterns impact soil aggregates and organic carbon of an alpine wetland on the Qinghai-Tibetan Plateau. Catena 244, 108249 (2024).Zhang, Y., Huang, X., Zhang, Z., Blewett, J. & Naafs, B. D. A. Spatiotemporal dynamics of dissolved organic carbon in a subtropical wetland and their implications for methane emissions. Geoderma 419, 115876 (2022).
    Google Scholar 
    Zhao, W. et al. Combination of mineral protection and molecular characteristics rather than alone to govern soil organic carbon stability in Qinghai-Tibetan plateau wetlands. J. Environ. Manag. 344, 118757 (2023).
    Google Scholar 
    Ding, Y. et al. The contribution of wetland plant litter to soil carbon pool: Decomposition rates and priming effects. Environ. Res. 224, 115575 (2023).
    Google Scholar 
    Xu, H. W. et al. Variation in soil organic carbon stability and driving factors after vegetation restoration in different vegetation zones on the Loess Plateau, China. Soil Tillage Res. 204, 12 (2020).
    Google Scholar 
    Malmer, N. & Wallén, B. Input rates, decay losses and accumulation rates of carbon in bogs during the last millennium: internal processes and environmental changes. Holocene 14, 111–117 (2004).
    Google Scholar 
    Ofiti, N. O. E. et al. Climate warming and elevated CO2 alter peatland soil carbon sources and stability. Nat. Commun. 14, 7533 (2023).
    Google Scholar 
    Liu, S. et al. Sedimentary ancient DNA reveals a threat of warming-induced alpine habitat loss to Tibetan Plateau plant diversity. Nat. Commun. 12, 2995 (2021).
    Google Scholar 
    Zimmermann, H. H. et al. Sedimentary ancient DNA and pollen reveal the composition of plant organic matter in Late Quaternary permafrost sediments of the Buor Khaya Peninsula (north-eastern Siberia). Biogeosciences 14, 575–596 (2017).
    Google Scholar 
    Fowler, H. J., Blenkinsop, S., Green, A. & Davies, P. A. Precipitation extremes in 2023. Nat. Rev. Earth Environ. 5, 250–252 (2024).
    Google Scholar 
    Sims, J. R. & Haby, V. A. Simplified colorimetric determination of soil organic matter. Soil Sci. 112, 137–141 (1971).
    Google Scholar 
    Folk, R. L. & Ward, W. C. Brazos River bar: a study in the significance of grain size parameters. J. Sediment. Res. 27, 3–26 (1957).
    Google Scholar 
    Ezcurra, E. Precision and bias of carbon storage estimations in wetland and mangrove sediments. Sci. Adv. 10, eadl1079 (2024).
    Google Scholar 
    Reimer, P. J. et al. The IntCal20 Northern Hemisphere radiocarbon age calibration curve (0–55 cal kBP). Radiocarbon 62, 725–757 (2020).
    Google Scholar 
    Blaauw, M. & Christen, J. A. Flexible paleoclimate age-depth models using an autoregressive gamma process. Bayesian Anal. 6, 457–474 (2011).
    Google Scholar 
    Hodgkins, S. B. et al. Tropical peatland carbon storage linked to global latitudinal trends in peat recalcitrance. Nat. Commun. 9, 3640 (2018).
    Google Scholar 
    Kylander, M. E., Ampel, L., Wohlfarth, B. & Veres, D. High-resolution X-ray fluorescence core scanning analysis of Les Echets (France) sedimentary sequence: new insights from chemical proxies. J. Quat. Sci. 26, 109–117 (2011).
    Google Scholar 
    Glass, J. B. et al. Molybdenum geochemistry in a seasonally dysoxic Mo-limited lacustrine ecosystem. Geochim. et. Cosmochim. Acta 114, 204–219 (2013).
    Google Scholar 
    Tenenhaus, M., Vinzi, V. E., Chatelin, Y.-M. & Lauro, C. PLS path modeling. Comput. Stat. Data Anal. 48, 159–205 (2005).
    Google Scholar 
    Wu, B. Medium resolution land cover data of Qinghai-Tibet Plateau (1980–2020). National Tibetan Plateau Data Center. https://doi.org/10.11888/Terre.tpdc.300593 (2023).China Geological Survey. Regional geological map of Xizang (1:250,000). National Geological Data Center. https://www.ngac.cn/ (2005).Download referencesAcknowledgementsThis work was supported by National Key Research and Development Program of China (Grant No. 2024YFF0808700), the National Natural Science Foundation of China (42407285), the Second Tibetan Plateau Scientific Expedition and Research Program (STEP) (Grant No. 2019QZKK0304), and the Key Science and Technology Program of Xizang (Grant No. XZ202101ZD0011G). We thank the Geothermal Geological Brigade (Tibet Bureau of Geology and Mineral Resources) for drilling and sampling, and Dr. Gao Shaopeng (Institute of Tibetan Plateau Research, CAS) for his supervision of the XRF scanning. No specific permissions were required for geological sampling. We sincerely thank the editor and the three anonymous reviewers for their constructive comments and insightful suggestions, which greatly improved the quality of this manuscript.Author informationAuthor notesThese authors contributed equally: Youqing Yang, Xiaoping Wang.These authors jointly supervised this work: Jianqing Du, Xiaoyong Cui.Authors and AffiliationsCollege of Life Sciences, University of Chinese Academy of Sciences, Beijing, ChinaYouqing Yang, Zeyuan Wang, Danhong Chen, Chenhao Gao, Yanbin Hao & Xiaoyong CuiBeijing Yanshan Earth Critical Zone National Research Station, University of Chinese Academy of Sciences, Beijing, ChinaYouqing Yang, Jianqing Du, Zhixiang Niu, Yanbin Hao, Haishan Niu, Kai Xue, Xiaoyong Cui & Yanfen WangInternational Joint Research Laboratory for Global Change Ecology, School of Life Sciences, Henan University, Kaifeng, ChinaXiaoping WangNational Key Laboratory of Earth System Numerical Modeling and Application, Chinese Academy of Sciences, Beijing, ChinaJianqing Du, Yanbin Hao & Kai XueCollege of Resources and Environment, University of Chinese Academy of Sciences, Beijing, ChinaJianqing Du, Zhixiang Niu, Yitao Zhang, Liyuan Ma, Haijun Zhang, Qiang Liu, Yu Wu, Haishan Niu, Kai Xue & Yanfen WangState Key Laboratory of Lithospheric and Environmental Coevolution, Institute of Geology and Geophysics, Chinese Academy of Sciences, Beijing, ChinaHui ZhangState Key Laboratory of Tibetan Plateau Earth System, Environment and Resources, Chinese Academy of Sciences, Beijing, ChinaYanfen WangAuthorsYouqing YangView author publicationsSearch author on:PubMed Google ScholarXiaoping WangView author publicationsSearch author on:PubMed Google ScholarJianqing DuView author publicationsSearch author on:PubMed Google ScholarHui ZhangView author publicationsSearch author on:PubMed Google ScholarZhixiang NiuView author publicationsSearch author on:PubMed Google ScholarYitao ZhangView author publicationsSearch author on:PubMed Google ScholarZeyuan WangView author publicationsSearch author on:PubMed Google ScholarLiyuan MaView author publicationsSearch author on:PubMed Google ScholarHaijun ZhangView author publicationsSearch author on:PubMed Google ScholarDanhong ChenView author publicationsSearch author on:PubMed Google ScholarChenhao GaoView author publicationsSearch author on:PubMed Google ScholarQiang LiuView author publicationsSearch author on:PubMed Google ScholarYu WuView author publicationsSearch author on:PubMed Google ScholarYanbin HaoView author publicationsSearch author on:PubMed Google ScholarHaishan NiuView author publicationsSearch author on:PubMed Google ScholarKai XueView author publicationsSearch author on:PubMed Google ScholarXiaoyong CuiView author publicationsSearch author on:PubMed Google ScholarYanfen WangView author publicationsSearch author on:PubMed Google ScholarContributionsJ.D. and Yanfen Wang conceived the study and designed the research protocol. Y.Y., Z.N., Y.Z., Z.W., L.M., Q.L., and Yu Wu carried out the field investigation and sample collection. Y.Y., X.W., Haijun Zhang, D.C., and C.G. performed data processing and statistical analyses. Y.Y. and X.W. drafted the initial manuscript. J.D., Hui Zhang, Y.H., H.N., K.X., and X.C. contributed to data interpretation and critically revised the manuscript. Yanfen Wang acquired funding for the project.Corresponding authorsCorrespondence to
    Jianqing Du or Xiaoyong Cui.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Peer review

    Peer review information
    Communications Earth and Environment thanks Alberto Araneda and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Primary Handling Editors: Somaparna Ghosh [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 InformationReporting SummaryRights 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 articleYang, Y., Wang, X., Du, J. et al. Climate and sedimentary structure drive deep labile carbon accumulation in alpine wetlands.
    Commun Earth Environ (2025). https://doi.org/10.1038/s43247-025-03081-8Download citationReceived: 25 April 2025Accepted: 01 December 2025Published: 16 December 2025DOI: https://doi.org/10.1038/s43247-025-03081-8Share 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

    A general framework for nitrogen deposition effects on soil respiration in global forests

    AbstractSince the Industrial Revolution, human activities have altered atmospheric nitrogen (N) deposition to global forests, affecting carbon dioxide emissions from soils (soil respiration or SR) – one of the largest land-atmosphere carbon fluxes. However, experimental studies have demonstrated both positive and negative effects of N deposition on SR in global forests, leading to debates on how N deposition increases or decreases SR. We developed a framework for generalizing SR responses to N deposition using synthesized data from 168 N addition experiments worldwide and observed SR across the global natural N deposition gradient. The findings indicate that N deposition decreased SR in 2.9% of global forested areas, particularly in eastern China, western Europe, and the eastern USA. However, the net effect of N deposition increased the global forest SR by ~5% (1.7 ± 0.1 PgC yr–1). If N pollution could be effectively controlled, global forest SR would decrease, potentially contributing to a reduction in the terrestrial carbon emissions.

    Similar content being viewed by others

    Temporal patterns of soil carbon emission in tropical forests under long-term nitrogen deposition

    Article

    01 December 2022

    Unexpected sustained soil carbon flux in response to simultaneous warming and nitrogen enrichment compared with single factors alone

    Article

    24 September 2024

    Nitrogen deposition contributed to a global increase in nitrous oxide emissions from forest soils

    Article
    Open access
    28 September 2024

    Data availability

    Data supporting the findings of this study (including CO2_exp and CO2_obs datasets) are available in Zenodo (https://doi.org/10.5281/zenodo.17670031).
    Code availability

    R code file supporting the findings of this study is available in Zenodo (https://doi.org/10.5281/zenodo.17670031).
    ReferencesFriedlingstein, P. et al. Global Carbon Budget 2022. Earth Syst. Sci. Data 14, 4811–4900 (2022).
    Google Scholar 
    Lei, J. et al. Temporal changes in global soil respiration since 1987. Nat. Commun. 12, 403 (2021).
    Google Scholar 
    Lu, H. B. et al. Comparing machine learning-derived global estimates of soil respiration and its components with those from terrestrial ecosystem models. Environ. Res Lett. 16, 14 (2021).
    Google Scholar 
    Raich, J. W. & Schlesinger, W. H. The global carbon dioxide flux in soil respiration and its relationship to vegetation and climate. Tellus B 44, 81–99 (1992).
    Google Scholar 
    Cai, W. X. et al. Carbon sequestration of Chinese forests from 2010 to 2060 spatiotemporal dynamics and its regulatory strategies. Sci. Bull. 67, 836–843 (2022).
    Google Scholar 
    Mo, L. et al. Integrated global assessment of the natural forest carbon potential. Nature 624, 92–101 (2023).
    Google Scholar 
    Griscom, B. W. et al. Natural climate solutions. Proc. Natl. Acad. Sci. USA 114, 11645–11650 (2017).
    Google Scholar 
    Galloway, J. N. et al. Nitrogen cycles: Past, present, and future. Biogeochemistry 70, 153–226 (2004).
    Google Scholar 
    Ackerman, D., Millet, D. B. & Chen, X. Global estimates of inorganic nitrogen deposition across four decades. Glob. Biogeochem. Cycles 33, 100–107 (2019).
    Google Scholar 
    Liu, L., Wen, Z., Liu, S., Zhang, X. & Liu, X. Decline in atmospheric nitrogen deposition in China between 2010 and 2020. Nat. Geosci. 17, 733–736 (2024).
    Google Scholar 
    Aber, J. D. et al. Nitrogen saturation in temperate forest ecosystems: hypotheses revisited. BioScience 48, 921–934 (1998).
    Google Scholar 
    Fenn, M. E. et al. Nitrogen excess in North American ecosystems: predisposing factors, ecosystem responses, and management strategies. Ecol. Appl 8, 706–733 (1998).
    Google Scholar 
    Cen, X. et al. Global patterns of nitrogen saturation in forests. One Earth 8, 101132 (2025).
    Google Scholar 
    Du, E. et al. Global patterns of terrestrial nitrogen and phosphorus limitation. Nat. Geosci. 13, 221–226 (2020).
    Google Scholar 
    Cen, X. et al. Suppression of nitrogen deposition on global forest soil CH4 uptake depends on nitrogen status. Glob. Biogeochem. Cycles 38, e2024GB008098 (2024).
    Google Scholar 
    Cleveland, C. C. & Townsend, A. R. Nutrient additions to a tropical rain forest drive substantial soil carbon dioxide losses to the atmosphere. Proc. Natl. Acad. Sci. USA 103, 10316–10321 (2006).
    Google Scholar 
    Mo, J. et al. Nitrogen addition reduces soil respiration in a mature tropical forest in southern China. Glob. Change Biol. 14, 403–412 (2008).
    Google Scholar 
    Bowden, R. D., Davidson, E., Savage, K., Arabia, C. & Steudler, P. Chronic nitrogen additions reduce total soil respiration and microbial respiration in temperate forest soils at the Harvard Forest. For. Ecol. Manag. 196, 43–56 (2004).
    Google Scholar 
    Janssens, I. A. et al. Reduction of forest soil respiration in response to nitrogen deposition. Nat. Geosci. 3, 315–322 (2010).
    Google Scholar 
    Liu, Y. et al. Spatially explicit estimate of nitrogen effects on soil respiration across the globe. Glob. Change Biol. 29, 3591 (2023).
    Google Scholar 
    Chen, C. & Chen, H. Y. H. Mapping global nitrogen deposition impacts on soil respiration. Sci. Total Environ. 871, 161986 (2023).
    Google Scholar 
    Bond-Lamberty, B. et al. Twenty Years of Progress, Challenges, and Opportunities in Measuring and Understanding Soil Respiration. J. Geophys. Res.: Biogeosciences 129, e2023JG007637 (2024).
    Google Scholar 
    Zhou, L. Y. et al. Different responses of soil respiration and its components to nitrogen addition among biomes: a meta-analysis. Glob. Change Biol. 20, 2332–2343 (2014).
    Google Scholar 
    de Vries, W., Du, E. Z. & Butterbach-Bahl, K. Short and long-term impacts of nitrogen deposition on carbon sequestration by forest ecosystems. Curr. Opin. Environ. Sustainability 9-10, 90–104 (2014).
    Google Scholar 
    Li, X. Y. et al. The contrasting effects of deposited NH4+ and NO3- on soil CO2, CH4 and N2O fluxes in a subtropical plantation, southern China. Ecol. Eng. 85, 317–327 (2015).
    Google Scholar 
    Bai, Y. et al. Tradeoffs and thresholds in the effects of nitrogen addition on biodiversity and ecosystem functioning: evidence from inner Mongolia Grasslands. Glob. Change Biol. 16, 358–372 (2010).
    Google Scholar 
    Reich, P. B., Tjoelker, M. G., Machado, J.-L. & Oleksyn, J. Universal scaling of respiratory metabolism, size and nitrogen in plants. Nature 439, 457–461 (2006).
    Google Scholar 
    Egidi, E., Coleine, C., Delgado-Baquerizo, M. & Singh, B. K. Assessing critical thresholds in terrestrial microbiomes. Nat. Microbiol. 8, 2230–2233 (2023).
    Google Scholar 
    Michaelis, L. & Menten, M. L. Die kinetik der invertinwirkung. Biochem. z. 49, 352 (1913).
    Google Scholar 
    Zhou, Z., Wang, C., Zheng, M., Jiang, L. & Luo, Y. Patterns and mechanisms of responses by soil microbial communities to nitrogen addition. Soil Biol. Biochem. 115, 433–441 (2017).
    Google Scholar 
    LeBauer, D. S. & Treseder, K. K. Nitrogen limitation of net primary productivity in terrestrial ecosystems is globally distributed. Ecology 89, 371–379 (2008).
    Google Scholar 
    Geisseler, D. & Horwath, W. R. Relationship between carbon and nitrogen availability and extracellular enzyme activities in soil. Pedobiologia 53, 87–98 (2009).
    Google Scholar 
    Allison, S. D. & Vitousek, P. M. Responses of extracellular enzymes to simple and complex nutrient inputs. Soil Biol. Biochem. 37, 937–944 (2005).
    Google Scholar 
    Goyal, S. S. & Huffaker, R. C. In Nitrogen in Crop Production (ed. Hauck, R. D.) (ASA, CSSA, SSSA;) 97–118 (1984).Kreutzer, K., Butterbach-Bahl, K., Rennenberg, H. & Papen, H. The complete nitrogen cycle of an N-saturated spruce forest ecosystem. Plant Biol. 11, 643–649 (2009).
    Google Scholar 
    Costantini, D., Metcalfe, N. B. & Monaghan, P. Ecological processes in a hormetic framework. Ecol. Lett. 13, 1435–1447 (2010).
    Google Scholar 
    Morgan Ernest, S. K. & Brown, J. H. Homeostasis and compensation: The role of species and resources in ecosystem stability. Ecology 82, 2118–2132 (2001).
    Google Scholar 
    Gilliam, F. S. Response of the herbaceous layer of forest ecosystems to excess nitrogen deposition. J. Ecol. 94, 1176–1191 (2006).
    Google Scholar 
    Bobbink, R. et al. Global assessment of nitrogen deposition effects on terrestrial plant diversity: a synthesis. Ecol. Appl. 20, 30–59 (2010).
    Google Scholar 
    Zheng, M. et al. Temporal patterns of soil carbon emission in tropical forests under long-term nitrogen deposition. Nat. Geosci. 15, 1002–1010 (2022).
    Google Scholar 
    Jian J. et al. A restructured and updated global soil respiration database (SRDB-V5). Earth Syst. Sci. Data 13, 255–267 (2021).Fog, K. The effect of added nitrogen on the rate of decomposition of organic matter. Biol. Rev. 63, 433–462 (1988).
    Google Scholar 
    Erofeeva, E. A. Environmental hormesis: from cell to ecosystem. Curr. Opin. Environ. Sci. Health 29, 100378 (2022).
    Google Scholar 
    Amyntas, A. et al. Niche complementarity among plants and animals can alter the biodiversity–ecosystem functioning relationship. Funct. Ecol. 37, 2652–2665 (2023).
    Google Scholar 
    Band, N., Kadmon, R., Mandel, M. & DeMalach, N. Assessing the roles of nitrogen, biomass, and niche dimensionality as drivers of species loss in grassland communities. Proc. Natl. Acad. Sci. 119, e2112010119 (2022).
    Google Scholar 
    Wang, C. et al. Long-term nitrogen input reduces soil bacterial network complexity by shifts in life history strategy in temperate grassland. iMeta 3, e194 (2024).
    Google Scholar 
    Liu, J. et al. Nitrogen addition reduced ecosystem stability regardless of its impacts on plant diversity. J. Ecol. 107, 2427–2435 (2019).
    Google Scholar 
    Melillo, J. M. et al. Long-term pattern and magnitude of soil carbon feedback to the climate system in a warming world. Science 358, 101–104 (2017).
    Google Scholar 
    Levin, S. A. The problem of pattern and scale in ecology: the Robert H. MacArthur award lecture. Ecology 73, 1943–1967 (1992).
    Google Scholar 
    Bae, K., Fahey, T. J., Yanai, R. D. & Fisk, M. Soil nitrogen availability affects belowground carbon allocation and soil respiration in northern hardwood forests of new hampshire. Ecosystems 18, 1179–1191 (2015).
    Google Scholar 
    Haynes, B. E. & Gower, S. T. Belowground carbon allocation in unfertilized and fertilized red pine plantations in Northern Wisconsin. Tree Physiol. 15, 317–325 (1995).
    Google Scholar 
    Wallenstein, M. D., McNulty, S., Fernandez, I. J., Boggs, J. & Schlesinger, W. H. Nitrogen fertilization decreases forest soil fungal and bacterial biomass in three long-term experiments. For. Ecol. Manag. 222, 459–468 (2006).
    Google Scholar 
    Treseder, K. K. Nitrogen additions and microbial biomass: a meta-analysis of ecosystem studies. Ecol. Lett. 11, 1111–1120 (2008).
    Google Scholar 
    Zhang, T. A., Chen, H. Y. H. & Ruan, H. Global negative effects of nitrogen deposition on soil microbes. ISME J. 12, 1817–1825 (2018).
    Google Scholar 
    Xing, A. J. et al. Nonlinear responses of ecosystem carbon fluxes to nitrogen deposition in an old-growth boreal forest. Ecol. Lett. 25, 77–88 (2022).
    Google Scholar 
    Zhang, D. et al. Changes in above-/below-ground biodiversity and plant functional composition mediate soil respiration response to nitrogen input. Funct. Ecol. 35, 1171–1182 (2021).
    Google Scholar 
    Lamanna, C. et al. Functional trait space and the latitudinal diversity gradient. Proc. Natl Acad. Sci. 111, 13745–13750 (2014).
    Google Scholar 
    Hillebrand, H. On the Generality of the Latitudinal Diversity Gradient. Am. Naturalist 163, 192–211 (2004).
    Google Scholar 
    Du, E. & de Vries, W. Links between nitrogen limitation and saturation in terrestrial ecosystems. Glob. Change Biol. 31, e70271 (2025).
    Google Scholar 
    Bond-Lamberty, B., Bronson, D., Bladyka, E. & Gower, S. T. A comparison of trenched plot techniques for partitioning soil respiration. Soil Biol. Biochem. 43, 2108–2114 (2011).
    Google Scholar 
    Wang, W., Chen, W. & Wang, S. Forest soil respiration and its heterotrophic and autotrophic components: Global patterns and responses to temperature and precipitation. Soil Biol. Biochem. 42, 1236–1244 (2010).
    Google Scholar 
    Zhao, Z., Ding, X., Wang, G. & Li, Y. 30 m Resolution Global Maps of Forest Soil Respiration and Its Changes From 2000 to 2020. Earth’s Future 12, e2023EF004007 (2024).
    Google Scholar 
    Cottingham, K. L., Lennon, J. T. & Brown, B. L. Knowing when to draw the line: designing more informative ecological experiments. Front. Ecol. Environ. 3, 145–152 (2005).
    Google Scholar 
    Yu, G. et al. Stabilization of atmospheric nitrogen deposition in China over the past decade. Nat. Geosci. 12, 424–429 (2019).
    Google Scholar 
    Liu, Y. et al. Spatially explicit estimate of nitrogen effects on soil respiration across the globe. Glob. Change Biol. 29, 3591–3600 (2023).
    Google Scholar 
    Yan, W., Zhong, Y., Yang, J., Shangguan, Z. & Torn, M. S. Response of soil greenhouse gas fluxes to warming: A global meta-analysis of field studies. Geoderma 419, 115865 (2022).
    Google Scholar 
    Shangguan, W., Dai, Y., Duan, Q., Liu, B. & Yuan, H. A global soil data set for earth system modeling. J. Adv. Modeling Earth Syst. 6, 249–263 (2014).
    Google Scholar 
    Liu, H. et al. Annual dynamics of global land cover and its long-term changes from 1982 to 2015. Earth Syst. Sci. Data 12, 1217–1243 (2020).
    Google Scholar 
    Hansen, M. C., Stehman, S. V. & Potapov, P. V. Quantification of global gross forest cover loss. Proc. Natl. Acad. Sci. 107, 8650–8655 (2010).
    Google Scholar 
    Wang, C. & Tang, Y. Responses of plant phenology to nitrogen addition: a meta-analysis. Oikos 128, 1243–1253 (2019).
    Google Scholar 
    Xu, C. et al. Long-term, amplified responses of soil organic carbon to nitrogen addition worldwide. Glob. Change Biol. 27, 1170–1180 (2021).
    Google Scholar 
    Wang, Z., Xing, A. & Shen, H. Effects of nitrogen addition on the combined global warming potential of three major soil greenhouse gases: A global meta-analysis. Environ. Pollut. 334, 121848 (2023).
    Google Scholar 
    R Core Team. R: A Language and Environment For Statistical Computing (R Foundation for Statistical Computing, 2020).Breiman, L. Random Forests. Mach. Learn. 45, 5–32 (2001).
    Google Scholar 
    Bond-Lamberty, B. & Thomson, A. Temperature-associated increases in the global soil respiration record. Nature 464, 579–U132 (2010).
    Google Scholar 
    Reay, D. S., Dentener, F., Smith, P., Grace, J. & Feely, R. A. Global nitrogen deposition and carbon sinks. Nat. Geosci. 1, 430–437 (2008).
    Google Scholar 
    Wang, W. J., Dalal, R. C., Moody, P. W. & Smith, C. J. Relationships of soil respiration to microbial biomass, substrate availability and clay content. Soil Biol. Biochem. 35, 273–284 (2003).
    Google Scholar 
    Chen, S., Zou, J., Hu, Z., Chen, H. & Lu, Y. Global annual soil respiration in relation to climate, soil properties and vegetation characteristics: Summary of available data. Agr. For. Meteorol. 198-199, 335–346 (2014).
    Google Scholar 
    Huang, N. et al. Spatial and temporal variations in global soil respiration and their relationships with climate and land cover. Sci. Adv. 6, 11 (2020).
    Google Scholar 
    Liaw, A. & Wiener, M. Classification and regression by randomForest. R. N. 2, 18–22 (2002).
    Google Scholar 
    Muggeo, V. M. Segmented: an R package to fit regression models with broken-line relationships. R. N. 8, 20–25 (2008).
    Google Scholar 
    Ritz, C., Baty, F., Streibig, J. C. & Gerhard, D. Dose-response analysis using R. PLoS ONE 10, e0146021 (2016).
    Google Scholar 
    ESRI. ArcGIS Desktop: Release 10 (ESRI, 2011).Mason, R. E. et al. Evidence, causes, and consequences of declining nitrogen availability in terrestrial ecosystems. Science 376, eabh3767 (2022).
    Google Scholar 
    Download referencesAcknowledgementsThe authors thank Prof. James Raich (Iowa State University) for his inputs on early version of this paper. We are grateful to all the researchers who spent time and effort on manipulative experiments and observations and made the data publicly accessible. This work was financially supported by the National Natural Science Foundation of China (32430067, 32588202, 42141004) and the National Key R&D Program of China (2023YFF1305900, 2022YFF080210102) received by N.H., and the Pioneer Center for Landscape Research in Sustainable Agricultural Futures (Land-CRAFT), DNRF grant number P2 received by K.B.B.Author informationAuthors and AffiliationsKey Laboratory of Boreal Forest Ecosystem Conservation and Restoration, National Forestry and Grassland Administration, Harbin, ChinaXiaoyu Cen & Nianpeng HeKey Laboratory of Ecosystem Network Observation and Modeling, Institute of Geographic Sciences and Natural Resources Research, Chinese Academy of Sciences, Beijing, ChinaXiaoyu Cen, Nianpeng He, Shuli Niu, Kailiang Yu, Li Xu & Mingxu LiDepartment of Earth System Science, Stanford University, Stanford, CA, USAXiaoyu Cen, Peter Vitousek & Elizabeth L. PaulusPioneer Center Land-CRAFT, Department of Agroecology, Aarhus University, Aarhus C, DenmarkXiaoyu Cen & Klaus Butterbach-BahlInstitute of Carbon Neutrality, School of Ecology, Northeast Forestry University, Harbin, ChinaNianpeng HeJoint Global Change Research Institute, Pacific Northwest National Laboratory, College Park, MD, USABen Bond-LambertyState Key Laboratory of Earth Surface Processes and Disaster Risk Reduction, Faculty of Geographical Science, Beijing Normal University, Beijing, ChinaEnzai DuHigh Meadows Environmental Institute, Princeton University, Princeton, NJ, USAKailiang YuKey Laboratory of Vegetation Restoration and Management of Degraded Ecosystems, Guangdong Provincial Key Laboratory of Applied Botany, South China Botanical Garden, Chinese Academy of Sciences, Guangzhou, ChinaMianhai ZhengDepartment of System Earth Science, Maastricht University, Innovalaan 1, Venlo, the NetherlandsKevin Van SundertGeochemistry and Biogeochemistry Group, SLAC National Accelerator Laboratory, Menlo Park, CA, USAElizabeth L. PaulusNicholas School of the Environment, Duke University, Durham, NC, USALiyin HeInstitute for Meteorology and Climate Research, Atmospheric Environmental Research, Karlsruhe Institute of Technology, Garmisch-Partenkirchen, GermanyKlaus Butterbach-BahlAuthorsXiaoyu CenView author publicationsSearch author on:PubMed Google ScholarPeter VitousekView author publicationsSearch author on:PubMed Google ScholarNianpeng HeView author publicationsSearch author on:PubMed Google ScholarBen Bond-LambertyView author publicationsSearch author on:PubMed Google ScholarShuli NiuView author publicationsSearch author on:PubMed Google ScholarEnzai DuView author publicationsSearch author on:PubMed Google ScholarKailiang YuView author publicationsSearch author on:PubMed Google ScholarMianhai ZhengView author publicationsSearch author on:PubMed Google ScholarKevin Van SundertView author publicationsSearch author on:PubMed Google ScholarElizabeth L. PaulusView author publicationsSearch author on:PubMed Google ScholarLiyin HeView author publicationsSearch author on:PubMed Google ScholarLi XuView author publicationsSearch author on:PubMed Google ScholarMingxu LiView author publicationsSearch author on:PubMed Google ScholarKlaus Butterbach-BahlView author publicationsSearch author on:PubMed Google ScholarContributionsX.C. collected data and carried out the analysis based on feedback from N.H., P.V., and K.B.B. X.C. drafted the initial manuscript, P.V., N.H., B.B.L., S.N., E.D., K.Y., M.Z., K.V.S., E.L.P., L.H., L.X., M.L., and K.B.B. reviewed and edited the manuscript.Corresponding authorCorrespondence to
    Nianpeng He.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 fileRights 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 articleCen, X., Vitousek, P., He, N. et al. A general framework for nitrogen deposition effects on soil respiration in global forests.
    Nat Commun (2025). https://doi.org/10.1038/s41467-025-67203-8Download citationReceived: 07 February 2025Accepted: 25 November 2025Published: 16 December 2025DOI: https://doi.org/10.1038/s41467-025-67203-8Share 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

    Social environment affects vocal individuality in a non-learning species

    AbstractIndividual recognition is fundamental to the social behaviour of many animals. In the context of territorial behaviour, animals in high-density populations encounter conspecific rivals and potential mates more frequently, which should enhance the individuality of territorial signals to facilitate recognition among conspecifics. We investigated vocal individuality in male territorial calls of two populations of little owls (Athene noctua) with different densities. Further, to explore the potential influence of local population distribution on individuality, we also examined isolated males without neighbours and clumped males with neighbours. Our findings indicate higher individuality at higher densities across both scenarios, measured using two individuality metrics: Beecher’s information statistic and Discrimination score. Clumped males exhibited significantly lower acoustic niche overlaps (i.e. higher vocal individuality) compared to isolated males. However, only a non-significant trend for lower acoustic niche overlaps (i.e. higher vocal individuality) was found for males from high density compared to low density populations. This suggests that the immediate social environment might be more influential than larger-scale population density patterns. This study suggests that vocal individuality in a territorial species is influenced by conspecific density, similar to findings in group-living and colonial species.

    Similar content being viewed by others

    Species-independent analysis and identification of emotional animal vocalizations

    Article
    Open access
    06 August 2025

    Density-dependent network structuring within and across wild animal systems

    Article

    04 September 2025

    Identifying unknown Indian wolves by their distinctive howls: its potential as a non-invasive survey method

    Article
    Open access
    31 March 2021

    Data availability

    Data available on figshare: [doi.org/10.6084/m9.figshare.29958992](http:/doi.org/10.6084/m9.figshare.29958992).
    ReferencesJudge, P. G. & de Waal, F. B. Rhesus monkey behaviour under diverse population densities: coping with long-term crowding. Anim. Behav. 54, 643–662 (1997).
    Google Scholar 
    Jirotkul, M. Population density influences male-male competition in guppies. Anim. Behav. 58, 1169–1175 (1999).
    Google Scholar 
    Knell, R. J. Population density and the evolution of male aggression. J. Zool. 278, 83–90 (2009).
    Google Scholar 
    Cooper, W. E., Dimopoulos, I., Pafilis, P. & Sex Age, and population density affect aggressive behaviors in island lizards promoting cannibalism. Ethology 121, 260–269 (2015).
    Google Scholar 
    Luna, Á., Palma, A., Sanz-Aguilar, A., Tella, J. L. & Carrete, M. Personality-dependent breeding dispersal in rural but not urban Burrowing owls. Sci. Rep. 9, 2886 (2019).
    Google Scholar 
    Fedy, B. C. & Stutchbury, B. J. Territory switching and floating in White-bellied antbird (Myrmeciza longipes), a resident tropical passerine in Panama. Auk 121, 486–496 (2004).
    Google Scholar 
    Schoepf, I., Schmohl, G., König, B., Pillay, N. & Schradin, C. Manipulation of population density and food availability affects home range sizes of African striped mouse females. Anim. Behav. 99, 53–60 (2015).
    Google Scholar 
    Arcese, P. & Smith, J. N. Effects of population density and supplemental food on reproduction in Song sparrows. J. Anim. Ecol. 57, 119–136 (1988).
    Google Scholar 
    Sofaer, H. R., Sillett, T. S., Langin, K. M., Morrison, S. A. & Ghalambor, C. K. Partitioning the sources of demographic variation reveals density-dependent nest predation in an Island bird population. Ecol. Evol. 4, 2738–2748 (2014).
    Google Scholar 
    Emlen, S. T. & Oring, L. W. Ecology, sexual selection, and the evolution of mating systems. Science 197, 215–223 (1977).
    Google Scholar 
    Lukas, D. & Clutton-Brock, T. H. The evolution of social monogamy in mammals. Science 341, 526–530 (2013).
    Google Scholar 
    Morales, M. B. et al. Density dependence and habitat quality modulate the intensity of display territory defence in an exploded lekking species. Behav. Ecol. Sociobiol. 68, 1493–1504 (2014).
    Google Scholar 
    Beecher, M. D. Signalling systems for individual recognition: an information theory approach. Anim. Behav. 38, 248–261 (1989).
    Google Scholar 
    Pollard, K. A., Blumstein, D. T. & Griffin, S. C. Pre-screening acoustic and other natural signatures for use in noninvasive individual identification. J. Appl. Ecol. 47, 1103–1109 (2010).
    Google Scholar 
    Freeberg, T. M. Social complexity can drive vocal complexity: group size influences vocal information in Carolina chickadees. Psychol. Sci. 17, 557–561 (2006).
    Google Scholar 
    Freeberg, T. M., Dunbar, R. I. & Ord, T. J. Social complexity as a proximate and ultimate factor in communicative complexity. Philosophical Trans. Royal Soc. B: Biol. Sci. 367, 1785–1801 (2012).
    Google Scholar 
    Peckre, L., Kappeler, P. M. & Fichtel, C. Clarifying and expanding the social complexity hypothesis for communicative complexity. Behav. Ecol. Sociobiol. 73, 1–19 (2019).
    Google Scholar 
    Arnold, B. D. & Wilkinson, G. S. Individual specific contact calls of Pallid bats (Antrozous pallidus) attract conspecifics at roosting sites. Behav. Ecol. Sociobiol. 65, 1581–1593 (2011).
    Google Scholar 
    Charrier, I., Aubin, T. & Mathevon, N. Mother-calf vocal communication in Atlantic walrus: a first field experimental study. Anim. Cogn. 13, 471–482 (2010).
    Google Scholar 
    Pitcher, B. J., Harcourt, R. G. & Charrier, I. Rapid onset of maternal vocal recognition in a colonially breeding mammal, the Australian sea lion. PLoS One 5(8), e12195 (2010).
    Google Scholar 
    Sheehan, M. J. et al. Selection on coding and regulatory variation maintains individuality in major urinary protein scent marks in wild mice. PLoS Genet. 12(1), e1005891 (2016).
    Google Scholar 
    Tibbetts, E. A. Complex social behaviour can select for variability in visual features: a case study in Polistes wasps. Proc. R Soc. Lond. B Biol. Sci. 271, 1955–1960 (2004).
    Google Scholar 
    Beecher, M. D., Medvin, M. B., Stoddard, P. K. & Loesche, P. Acoustic adaptations for parent-offspring recognition in swallows. Exp. Biol. 45, 179–193 (1986).
    Google Scholar 
    Loesche, P., Stoddard, P. K., Higgins, B. J. & Beecher, M. D. Signature versus perceptual adaptations for individual vocal recognition in swallows. Behaviour 118, 15–25 (1991).
    Google Scholar 
    Medvin, M. B., Stoddard, P. K. & Beecher, M. D. Signals for parent-offspring recognition: a comparative analysis of the begging calls of Cliff swallows and Barn swallows. Anim. Behav. 45, 841–850 (1993).
    Google Scholar 
    Martin, M., Gridley, T., Elwen, S. H. & Charrier, I. Extreme ecological constraints lead to high degree of individual stereotypy in the vocal repertoire of the Cape fur seal (Arctocephalus pusillus pusillus). Behav. Ecol. Sociobiol. 75, 1–16 (2021).
    Google Scholar 
    Pollard, K. A. & Blumstein, D. T. Social group size predicts the evolution of individuality. Curr. Biol. 21, 413–417 (2011).
    Google Scholar 
    Wilkinson, G. S. Social and vocal complexity in bats. In Animal Social Complexity: Intelligence, Culture, and Individualized Societies (Harvard University Press, 2003). 
    Google Scholar 
    Brooks, R. J. & Falls, J. B. Individual recognition by song in White-throated sparrows. I. Discrimination of songs of neighbors and strangers. Can. J. Zool. 53, 879–888 (1975).
    Google Scholar 
    Godard, R. Long-term memory of individual neighbours in a migratory songbird. Nature 350, 228–229 (1991).
    Google Scholar 
    Hyman, J. & Hughes, M. Territory owners discriminate between aggressive and nonaggressive neighbours. Anim. Behav. 72, 209–215 (2006).
    Google Scholar 
    Jaška, P., Linhart, P. & Fuchs, R. Neighbour recognition in two sister songbird species with a simple and complex repertoire-a playback study. J. Avian Biol. 46, 151–158 (2015).
    Google Scholar 
    Temeles, E. J. The role of neighbours in territorial systems: when are they ‘dear enemies’? Anim. Behav. 47, 339–350 (1994).
    Google Scholar 
    Fisher, J. B. Evolution and bird sociality. Evol. As Process., 71–83. (1954).Blumstein, D. T., Mcclain, D. R. & Jesus, C. Alarcón-Nieto, G. Breeding bird density does not drive vocal individuality. Curr. Zool. 58, 765–772 (2012).
    Google Scholar 
    Delgado, M. D. M. et al. Population characteristics may reduce the levels of individual call identity. PLoS One 8(10), e77557 (2013).
    Google Scholar 
    Chen, Z. & Wiens, J. J. The origins of acoustic communication in vertebrates. Nat. Commun. 11, 1–8 (2020).
    Google Scholar 
    Ramanankirahina, R., Joly, M., Scheumann, M. & Zimmermann, E. The role of acoustic signaling for spacing and group coordination in a nocturnal, pair-living primate, the Western woolly Lemur (Avahi occidentalis. Am. J. Phys. Anthropol. 159, 466–477 (2016).
    Google Scholar 
    Zhang, C. M., Sun, C. N. & Lucas, F. Acoustic signal dominance in the multimodal communication of a nocturnal mammal. Curr. Zool. (2021).
    Google Scholar 
    Linhart, P. & Šálek, M. The assessment of biases in the acoustic discrimination of individuals. PLoS One 12(5), e0177206 (2017).
    Google Scholar 
    Odom, K. J., Slaght, J. C. & Gutiérrez, R. J. Distinctiveness in the territorial calls of Great horned owls within and among years. J. Raptor Res. 47, 21–30 (2013).
    Google Scholar 
    Yee, S. A., Puan, C. L., Chang, P. K. & Azhar, B. Vocal individuality of Sunda scops-owl (Otus lempiji) in Peninsular Malaysia. J. Raptor Res. 50, 379–390 (2016).
    Google Scholar 
    Madhavan, M. & Linhart, P. Vocal individuality in owls: a taxon-wide review in the context of Tinbergen’s four questions. J. Ornithol. 166, 307–319 (2025).
    Google Scholar 
    Cavanagh, P. M. & Ritchison, G. Variation in the bounce and whinny songs of the Eastern Screech-Owl. Wilson Bull. 99, 620–627 (1987).
    Google Scholar 
    Grieco, F. Aggregation of Eurasian scops owls Otus scops breeding in Magpie Pica pica nests. Ardea 106, 177–191 (2018).
    Google Scholar 
    Nagy, C. M. & Rockwell, R. F. Identification of individual Eastern Screech-Owls megascops Asio via vocalization analysis. Bioacoustics 21, 127–140 (2012).
    Google Scholar 
    Dreiss, A. N., Ruppli, C. A. & Roulin, A. Individual vocal signatures in barn Owl nestlings: does individual recognition have an adaptive role in sibling vocal competition? J. Evol. Biol. 27, 63–75 (2014).
    Google Scholar 
    Van Nieuwenhuyse, D., Van Harxen, R., Johnson, D. H. & De Raedt, J. The Little Owl: Population Dynamics, Behavior and Management of Athene noctua (Cambridge University Press, 2023). 
    Google Scholar 
    Chrenková, M., Dobrý, M. & Šálek, M. Further evidence of large-scale population decline and range contraction of the Little Owl Athene noctua in central Europe. Folia Zool. 66, 106–116 (2017).
    Google Scholar 
    Šálek, M. et al. Scale-dependent habitat associations of a rapidly declining farmland predator, the Little Owl Athene noctua, in contrasting agricultural landscapes. Agric. Ecosyst. Environ. 224, 56–66 (2016).
    Google Scholar 
    Le Gouar, P. J. et al. Long-term trends in survival of a declining population: the case of the Little Owl (Athene noctua) in the Netherlands. Oecologia 166, 369–379 (2011).
    Google Scholar 
    Żmihorski, M., Altenburg, D., Romanowski, J., Kowalski, M. & Osojca, G. Long term decline of the Little Owl (Athene noctua Scop., 1769) in central Poland. Pol. J. Ecol. 54, 321–324 (2006).
    Google Scholar 
    Mayer, M. et al. Fine-scale movement patterns and habitat selection of Little owls (Athene noctua) from two declining populations. PLoS One 16(9), e0256608 (2021).
    Google Scholar 
    Šálek, M. & Lövy, M. Spatial ecology and habitat selection of Little Owl Athene noctua during the breeding season in central European farmland. Bird. Conserv. Int. 22, 328–338 (2012).
    Google Scholar 
    Šálek, M. & Mayer, M. Farmstead modernization adversely affects farmland birds. J. Appl. Ecol. 60, 101–110 (2023).
    Google Scholar 
    Šálek, M. & Schröpfer, L. Population decline of the Little Owl (Athene noctua Scop.) in the Czech Republic. Pol. J. Ecol. 56, 527–534 (2008).
    Google Scholar 
    Šálek, M., Chrenkova, M. & Kipson, M. High population density of Little Owl (Athene noctua) in hortobagy National park, Hungary, central Europe. Pol. J. Ecol. 61, 165–169 (2013).
    Google Scholar 
    Šálek, M. et al. In Owl’s Paradise: Little Owl Population Densities in Traditional Human Settlements Represent One of the Highest Densities Reported among Owls. J. Raptor Res. 59(1), 1–11 (2025).
    Google Scholar 
    Jacobsen, L. B., Sunde, P., Rahbek, C., Dablesteen, T. & Thorup, K. Territorial calls in the Little Owl (Athene noctua): spatial dispersion and social interplay of mates and neighbours. Ornis Fenn. 90(1), 41–49 (2013).
    Google Scholar 
    Orlando, G., Varesio, A. & Chamberlain, D. Field evaluation for playback surveys: species-specific detection probabilities and distance estimation errors in a nocturnal bird community. Bird. Study. 68, 78–87 (2021).
    Google Scholar 
    Hardouin, L. A., Tabel, P. & Bretagnolle, V. Neighbour-stranger discrimination in the Little owl, Athene noctua. Anim. Behav. 72, 105–112 (2006).
    Google Scholar 
    Siracusa, E. R. et al. Familiar Neighbors, but not relatives, enhance fitness in a territorial mammal. Curr. Biol. 31, 438–445 (2021).
    Google Scholar 
    Falls, J. B. Individual recognition by sounds in birds. In Acoustic Communication in Birds Vol. 2 (eds Kroodsma, D. E. & Miller, E. H.) 237–278 (Academic, 1982).
    Google Scholar 
    Exo, K. M. Annual cycle and ecological adaptions in the Little Owl (Athene noctua). J. Ornithol. 129, 393–415 (1988).
    Google Scholar 
    Fick, S. E. & Hijmans, R. J. WorldClim 2: new 1-km Spatial resolution climate surfaces for global land areas. Intl J. Climatology. 37, 4302–4315 (2017).
    Google Scholar 
    Průchová, A., Šálek, M. & Linhart, P. Social factors affect vocal activity patterns of two common call types in Little Owl males. J. Ornithol. 166, 235–246 (2025).
    Google Scholar 
    Šálek, M. Dlouhodobý pokles početnosti sýčka obecného (Athene noctua) v jádrové oblasti Jeho rozšíření v Čechách. Sylvia 50, 2–12 (2014).
    Google Scholar 
    Exo. Population ecology of Little Owls Athene noctua in Central Europe: a review. The ecology and conservation of European owls, 64–75. (1992).Charif, R. A., Waack, A. M. & Strickman, L. M. Raven Pro 1.4 User’s Manual (Cornell Lab of Ornithology, 2010).Araya-Salas, M. Rraven: connecting R and Raven bioacoustic software. R package version 1.0.9. (2020).Linhart, P. et al. Measuring individual identity information in animal signals: overview and performance of available identity metrics. Methods Ecol. Evol. 10, 1558–1570 (2019).
    Google Scholar 
    Blumstein, D. T. & Munos, O. Individual, age and sex-specific information is contained in Yellow-bellied marmot alarm calls. Anim. Behav. 69, 353–361 (2005).
    Google Scholar 
    Tumulty, J. P., Lange, Z. K. & Bee, M. A. Identity signaling, identity reception, and the evolution of social recognition in a Neotropical frog. Evolution 76, 158–170 (2022).
    Google Scholar 
    Favaro, L., Gamba, M., Alfieri, C., Pessani, D. & McElligott, A. G. Vocal individuality cues in the African penguin (Spheniscus demersus): a source-filter theory approach. Sci. Rep. 5(1), 17255 (2015).
    Google Scholar 
    Galeotti, P., Paladin, M. & Pavan, G. Individually distinct hooting in male Pygmy owls Glaucidium passerinum: a multivariate approach. Ornis Scand. 1993, 15–20 (1993).
    Google Scholar 
    Li, Y., Xia, C., Lloyd, H., Li, D. & Zhang, Y. Identification of vocal individuality in male cuckoos using different analytical techniques. Avian Res. 8, 1–7 (2017).
    Google Scholar 
    Hutchinson, G. E. Concluding remarks. Cold Spring Harb Symp. Quant. Biol. 22, 415–427 (1957).
    Google Scholar 
    Alvarado-Serrano, D. F. & Knowles, L. L. Ecological niche models in phylogeographic studies: applications, advances and precautions. Mol. Ecol. Resour. 14, 233–248 (2014).
    Google Scholar 
    Bearhop, S., Adams, C. E., Waldron, S., Fuller, R. A. & MacLeod, H. Determining trophic niche width: a novel approach using stable isotope analysis. J. Anim. Ecol. 73, 1007–1012 (2004).
    Google Scholar 
    Raxworthy, C. J., Ingram, C. M., Rabibisoa, N. & Pearson, R. G. Applications of ecological niche modeling for species delimitation: a review and empirical evaluation using day geckos (Phelsuma) from Madagascar. Syst. Biol. 56, 907–923 (2007).
    Google Scholar 
    Hart, P. J., Ibanez, T., Paxton, K., Tredinnick, G., Sebastián-González, E., & Tanimoto-Johnson, A. Timing is everything: acoustic niche partitioning in two tropical wet forest bird communities. Front. Ecol. Evol. 9, 753363 (2021).Henry, C. S. & Wells, M. M. Acoustic niche partitioning in two cryptic sibling species of Chrysoperla green lacewings that must duet before mating. Anim. Behav. 80, 991–1003 (2010).
    Google Scholar 
    Sinsch, U., Lümkemann, K., Rosar, K., Schwarz, C. & Dehling, M. Acoustic niche partitioning in an Anuran community inhabiting an Afromontane wetland (Butare, Rwanda). Afr. Zool. 47, 60–73 (2012).
    Google Scholar 
    Swanson, H. K. et al. A new probabilistic method for quantifying n-dimensional ecological niches and niche overlap. Ecology 96, 318–324 (2015).
    Google Scholar 
    Core Team, R. R. C. R: A language and environment for statistical computing. (2022).Budka, M., Matyjasiak, P., Typiak, J., Okołowski, M. & Zagalska-Neubauer, M. Experienced males modify their behaviour during playback: the case of the Chaffinch. J. Ornithol. 160, 673–684 (2019).
    Google Scholar 
    Linhart, P., Fuchs, R., Poláková, S. & Slabbekoorn, H. Once bitten twice shy: long-term behavioural changes caused by trapping experience in Willow warblers Phylloscopus trochilus. J. Avian Biol. 43, 186–192 (2012).
    Google Scholar 
    Oñate-Casado, J., Porteš, M., Beran, V., Petrusek, A. & Petrusková, T. An experience to remember: lifelong effects of playback-based trapping on behaviour of a migratory passerine bird. Anim. Behav. 182, 19–29 (2021).
    Google Scholar 
    Sexton, K., Redmond, L., Murphy, M. & Dolan, A. Dawn song of Eastern kingbirds: intrapopulation variability and sociobiological correlates. Behav 144, 1273–1295 (2007).
    Google Scholar 
    Tobias, J. A., Gamarra-Toledo, V., García-Olaechea, D., Pulgarín, P. C. & Seddon, N. Year-round resource defence and the evolution of male and female song in suboscine birds: social armaments are mutual ornaments: evolution of mutual ornaments in birds. J. Evol. Biol. 24, 2118–2138 (2011).
    Google Scholar 
    Xia, C. et al. Dawn singing intensity of the male Brownish-Flanked Bush warbler: effects of territorial insertions and number of neighbors. Ethology 120, 324–330 (2014).
    Google Scholar 
    Ripmeester, E. A. P., Kok, J. S., Van Rijssel, J. C. & Slabbekoorn, H. Habitat-related birdsong divergence: a multi-level study on the influence of territory density and ambient noise in European Blackbirds. Behav. Ecol. Sociobiol. 64, 409–418 (2010).
    Google Scholar 
    Stuart, C. J., Grabarczyk, E. E., Vonhof, M. J. & Gill, S. A. Social factors, not anthropogenic noise or artificial light, influence onset of dawn singing in a common songbird. Auk 136, ukz045 (2019).
    Google Scholar 
    Owen, K. C. & Mennill, D. J. Singing in a fragmented landscape: Wrens in a tropical dry forest show sex differences in the effects of neighbours, time of day, and time of year. J. Ornithol. 162, 881–893 (2021).
    Google Scholar 
    Sánchez, N. V. & Mennill, D. J. Behavioural consequences of conspecific neighbours: a systematic literature review of the effects of local density on avian vocal communication. J. Ornithol. 165, 847–859 (2024).
    Google Scholar 
    Gokcekus, S., Firth, J. A., Regan, C. & Sheldon, B. C. Recognising the key role of individual recognition in social networks. Trends Ecol. Evol. 36, 1024–1035 (2021).
    Google Scholar 
    Tibbetts, E. A., Mullen, S. P. & Dale, J. Signal function drives phenotypic and genetic diversity: the effects of signalling individual identity, quality or behavioural strategy. Phil Trans. R Soc. B. 372, 20160347 (2017).
    Google Scholar 
    Osiecka, A. N., Briefer, E. F., Kidawa, D. & Wojczulanis-Jakubas, K. Strong individual distinctiveness across the vocal repertoire of a colonial seabird, the Little auk, Alle alle. Anim. Behav. 210, 199–211 (2024).
    Google Scholar 
    Charrier, I. Mother-offspring vocal recognition and social system in pinnipeds. Coding strategies in vertebrate acoustic communication 231–246 Cham: Springer International Publishing. (2020). Elfström, S. T. Responses of territorial Meadow pipits to strange and familiar song phrases in playback experiments. Anim. Behav. 40, 786–788 (1990).
    Google Scholar 
    Wegrzyn, E., Leniowski, K. & Osiejuk, T. S. Introduce yourself at the beginning-possible identification function of the initial part of the song in the Great reed warbler Acrocephalus arundinaceus. Ornis 86 (2), 61–70 (2009). (2009).
    Google Scholar 
    Mathevon, N. et al. Singing in the rain forest: how a tropical bird song transfers information. PLoS ONE. 3, e1580 (2008).
    Google Scholar 
    Wheeldon, A., Kwiatkowska, K., Szymański, P. & Osiejuk, T. S. Male and female songs propagation in a duetting tropical bird species in its preferred and secondary habitat. PLoS ONE. 17, e0275434 (2022).
    Google Scholar 
    Lengagne, T. Temporal stability in the individual features in the calls of Eagle owls (Bubo bubo). Behaviour 138, 1407–1419 (2001).
    Google Scholar 
    Garland, T. & Adolph, S. C. Why not to do two-Species comparative studies: limitations on inferring adaptation. Physiological Zool. 67, 797–828 (1994).
    Google Scholar 
    Goymann, W. & Schwabl, H. The tyranny of phylogeny—A plea for a less dogmatic stance on two-species comparisons: Funding bodies, journals and referees discourage two‐ or few‐species comparisons, but such studies provide essential insights complementary to phylogenetic comparative studies. BioEssays 43(8), 2100071 (2021).
    Google Scholar 
    Kidawa, D., Wojczulanis-Jakubas, K., Jakubas, D., Palme, R. & Barcikowski, M. Mine or my neighbours’ offspring: an experimental study on parental discrimination of offspring in a colonial seabird, the Little auk Alle alle. Sci. Rep. 13, 15088 (2023).
    Google Scholar 
    Klenova, A. V., Volodin, I. A. & Volodina, E. V. The variation in reliability of individual vocal signature throughout ontogenesis in the Red-crowned crane Grus japonensis. Acta ethol. 12, 29–36 (2009).
    Google Scholar 
    Lefevre, K., Montgomerie, R. & Gaston, A. J. Parent–offspring recognition in Thick-billed murres (Aves: Alcidae). Anim. Behav. 55, 925–938 (1998).
    Google Scholar 
    Osiecka, A. N., Oliva, M. Q., Kouřil, J., Petrusková, T. & Burchardt, L. S. Yellowhammer (Emberiza citrinella) males sing using individual isochronous rhythms and maximise rhythmic dissimilarity with neighbours. bioRxiv. Preprint at. https://doi.org/10.1101/2025.06.17.660106 (2025).
    Google Scholar 
    Geberzahn, N. & Aubin, T. How a songbird with a continuous singing style modulates its song when territorially challenged. Behav. Ecol. Sociobiol. 68, 1–12 (2014).
    Google Scholar 
    Hardouin, L. A., Reby, D., Bavoux, C., Burneleau, G. & Bretagnolle, V. Communication of male quality in owl hoots. Am. Nat. 169, 552–562 (2007).
    Google Scholar 
    Linhart, P., Jaška, P., Petrusková, T., Petrusek, A. & Fuchs, R. Being angry, singing fast? Signalling of aggressive motivation by syllable rate in a songbird with slow song. Behav. Processes. 100, 139–145 (2013).
    Google Scholar 
    McGregor, P. K., Dabelsteen, T., Shepherd, M. & Pedersen, S. B. The signal value of matched singing in Great tits: evidence from interactive playback experiments. Anim. Behav. 43, 987–998 (1992).
    Google Scholar 
    Thomsen, H. M., Balsby, T. J. & Dabelsteen, T. The imitation dilemma: can parrots maintain their vocal individuality when imitating conspecifics? Behaviour 156, 787–814 (2019).
    Google Scholar 
    Delport, W., Kemp, A. C. & Ferguson, J. W. H. Vocal identification of individual African wood owls Strix woodfordii: A technique to monitor long-term adult turnover and residency. Ibis 144, 30–39 (2002).
    Google Scholar 
    Rognan, C. B., Szewczak, J. M. & Morrison, M. L. Vocal individuality of Great gray owls in the Sierra Nevada. J. Wildl. Manage. 73, 755–760 (2009).
    Google Scholar 
    Takagi, M. Vocalizations of the Ryukyu scops owl Otus elegans: individually recognizable and stable. Bioacoustics 29, 28–44 (2020).
    Google Scholar 
    Tripp, T. M. & Otter, K. A. Vocal individuality as a potential long-term monitoring tool for Western screech-owls, Megascops kennicottii. Can. J. Zool. 84, 744–753 (2006).
    Google Scholar 
    Rivera-Gutierrez, H. F., Pinxten, R. & Eens, M. Songs differing in consistency elicit differential aggressive response in territorial birds. Biol. Lett. 7, 339–342 (2011).
    Google Scholar 
    Sierro, J., de Kort, S.R., Hartley, I.R. Sexual selection for both diversity and repetition in birdsong. Nat. Commun. 14(1), 3600 (2023).
    Google Scholar 
    Miles, L. S., Rivkin, L. R., Johnson, M. T., Munshi-South, J. & Verrelli, B. C. Gene flow and genetic drift in urban environments. Mol. Ecol. 28, 4138–4151 (2019).
    Google Scholar 
    Evans, K. L. Individual species and urbanisation. Urban ecology, 53-87 (2010).Hill, S. D., Aryal, A., Pawley, M. D. M. & Ji, W. So much for the city: Urban-rural song variation in a widespread Asiatic songbird. Integr. Zool. 13, 194–205 (2018).
    Google Scholar 
    Narango, D. L. & Rodewald, A. D. Urban-associated drivers of song variation along a rural–urban gradient. Behav. Ecol. 27, 608–616 (2016).
    Google Scholar 
    Halfwerk, W. et al. Adaptive changes in sexual signalling in response to urbanization. Nat. Ecol. Evol. 3, 374–380 (2018).
    Google Scholar 
    Brenowitz, E. A. Evolution of the vocal control system in the avian brain. Semin Neurosci. 3, 399–407 (1991).
    Google Scholar 
    Ten Cate, C. Re-evaluating vocal production learning in non-oscine birds. Phil Trans. R Soc. B 376(1836), 20200249 (2021).
    Google Scholar 
    Brumm, H. & Zollinger, S. A. The evolution of the Lombard effect: 100 years of psychoacoustic research. Behaviour 148, 1173–1198 (2011).
    Google Scholar 
    Dunlop, R. A., Cato, D. H. & Noad, M. J. Evidence of a Lombard response in migrating Humpback whales (Megaptera novaeangliae). J. Acoust. Soc. Am. 136, 430–437 (2014).
    Google Scholar 
    Halfwerk, W., Lea, A. M., Guerra, M. A., Page, R. A. & Ryan, M. J. Vocal responses to noise reveal the presence of the Lombard effect in a frog. Behav. Ecol. 27, 669–676 (2016).
    Google Scholar 
    Janik, V. M. & Slater, P. J. The different roles of social learning in vocal communication. Anim. Behav. 60, 1–11 (2000).
    Google Scholar 
    Arellano, C. M., Canelón, N. V., Delgado, S. & Berg, K. S. Allo-preening is linked to vocal signature development in a wild parrot. Behav. Ecol. 33, 202–212 (2022).
    Google Scholar 
    Beecher, M. & Brenowitz, E. Functional aspects of song learning in songbirds. Trends Ecol. Evol. 20, 143–149 (2005).
    Google Scholar 
    Palmero, A. M., Illera, J. C. & Laiolo, P. Song characterization in the Spectacled warbler (Sylvia conspicillata): a circum-Mediterranean species with a complex song structure. Bioacoustics 21, 175–191 (2012).
    Google Scholar 
    Slater, P. J. B. & Lachlan, R. F. Is Innovation in Bird Song Adaptive? In Animal Innovation (eds Reader, S. M. & Laland, K. N.) 117–136 (Oxford University Press, 2003).
    Google Scholar 
    Walcott, C., Mager, J. N. & Piper, W. Changing territories, changing tunes: male loons, Gavia immer, change their vocalizations when they change territories. Anim. Behav. 71, 673–683 (2006).
    Google Scholar 
    Groothuis, T. The influence of social experience on the development and fixation of the form of displays in the Black-headed gull. Anim. Behav. 43, 1–14 (1992).
    Google Scholar 
    Download referencesAcknowledgementsWe are grateful to all four anonymous reviewers whose feedback on a previous version of the manuscript helped to improve its quality and clarity.FundingThis study was supported by the Czech Science Foundation (GACR 21–04023 K). Martin Šálek was partly supported by the Czech Academy of Sciences in the framework of the program Strategy AV 21 and the research aim of the Czech Academy of Sciences (RVO 68081766). Malavika Madhavan was partly supported by the University of South Bohemia (GAJU No. 047/2025/P). Martin Šálek and Malavika Madhavan were also supported by One Nature Project (LIFE17 IPE/CZ/000005, LIFE IP: N2K Revisited), supported by the EU´s Financial Instrument LIFE.Author informationAuthors and AffiliationsDepartment of Zoology, Faculty of Science, University of South Bohemia, České Budějovice, Czech RepublicMalavika Madhavan, Lucie Hornátová, Alexandra Průchová & Pavel LinhartInstitute of Vertebrate Biology, Czech Academy of Sciences, Brno, Czech RepublicMartin ŠálekFaculty of Environmental Sciences, Czech University of Life Sciences Prague, Prague-Suchdol, Czech RepublicMartin ŠálekCenter for Sustainable Landscapes under Global Change (SustainScapes), Department of Biology, Aarhus University, Aarhus, DenmarkPavel LinhartAuthorsMalavika MadhavanView author publicationsSearch author on:PubMed Google ScholarLucie HornátováView author publicationsSearch author on:PubMed Google ScholarMartin ŠálekView author publicationsSearch author on:PubMed Google ScholarAlexandra PrůchováView author publicationsSearch author on:PubMed Google ScholarPavel LinhartView author publicationsSearch author on:PubMed Google ScholarContributionsPL and MŠ conceived the study and acquired funding for the project. MŠ and AP collected the data, and AP, PL, LH, and MM curated the data. MM, LH, and PL investigated the data and performed the analyses. MM and PL drafted the manuscript, and all authors contributed to its editing and final approval.Corresponding authorCorrespondence to
    Malavika Madhavan.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 1Supplementary Material 2Supplementary Material 3Rights 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 articleMadhavan, M., Hornátová, L., Šálek, M. et al. Social environment affects vocal individuality in a non-learning species.
    Sci Rep (2025). https://doi.org/10.1038/s41598-025-29387-3Download citationReceived: 23 July 2025Accepted: 17 November 2025Published: 15 December 2025DOI: https://doi.org/10.1038/s41598-025-29387-3Share 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
    KeywordsAcoustic nicheBioacousticsOwlPopulation densitySignal evolutionVocal learningVocal plasticity More

  • in

    Gender dynamics and remittances in the adoption of sustainable agricultural practices in Nepal

    AbstractAgricultural systems in rural Nepal face significant transformation due to climate change and shifting household labour dynamics. Male out-migration, a key but underexplored driver of this change, disrupts traditional gendered labor divisions and reshapes agricultural decision-making. In smallholder farms, traditionally, men are more responsible for tasks like ploughing and harvesting, whereas women take on planting, weeding, and winnowing roles. In the context of male out-migration, women must take primary responsibility for managing farms and households. However, persistent social and structural inequalities continue to constrain their decision-making authority. The feminisation of agriculture has important implications for adopting sustainable agricultural practices (SAPs). Using survey data from 400 households and Poisson regression analysis, this study examines the effects of migration, remittances, and female-managed farms on SAP adoption. Our results highlight that household farms having migrated member(s), receiving remittance and female-managed farms are more likely to adopt SAPs. In contrast, a higher number of out-migrating females negatively affects adoption, reflecting women’s critical role in sustainable farming adoption. Their participation in women’s groups, which provide training and financial resources and their management of tasks such as seed selection, winnowing, and organic pest control, are essential to SAP implementation. As such, our study provides a deeper understanding of the positive role of females in SAP adoption. We advocate for policies that recognising intersectional vulnerabilities, supporting women’s groups, lead to increased adoption of sustainable agricultural practices.

    Similar content being viewed by others

    Climate risks and adaptation strategies of farmers in East Africa and South Asia

    Article
    Open access
    18 May 2021

    Gender-based differences in eco-efficient farming

    Article
    Open access
    07 May 2025

    Empowering women in sustainable agriculture

    Article
    Open access
    26 March 2024

    Data availability

    The data that support the findings of this study are available from the corresponding author upon reasonable request. Due to ethical considerations and to protect participants’ privacy, the raw survey data cannot be made openly accessible.
    ReferencesLeder, S. Beyond the ‘Feminization of agriculture’: rural out-migration, shifting gender relations and emerging spaces in natural resource management. J. Rural Stud. 91, 157–169. https://doi.org/10.1016/j.jrurstud.2022.02.009 (2022).
    Google Scholar 
    Sunam, R., Sugden, F., Kharel, A., Sunuwar, T. R. & Ito, T. Unpacking youth engagement in agriculture: Land, labour mobility and youth livelihoods in rural Nepal. J. Agrarian Change. 25 (1), e12611. https://doi.org/10.1111/joac.12611 (2025).
    Google Scholar 
    Rayamajhee, V., Guo, W. & Bohara, A. K. The impact of climate change on rice production in Nepal. Econ. Disasters Clim. Change. 5, 111–134. https://doi.org/10.1007/s41885-020-00079-8 (2021).
    Google Scholar 
    CIAT; World Bank Climate-Smart agriculture in Nepal. CSA country profiles for Asia series. In International Centerfor Tropical Agriculture (CIAT); the World Bank; CGIAR Research Program on Climate Change, Agriculture and Food Security (CCAFS); Local Initiatives for Biodiversity Research and Development 26 (LI-BIRD), 2017).
    Google Scholar 
    Kandel, G. P., Bavorova, M., Ullah, A., Kaechele, H. & Pradhan, P. Building resilience to climate change: examining the impact of agro-ecological zones and social groups on sustainable development. Sustain. Dev. 31 (5), 3796–3810 (2023).
    Google Scholar 
    Gomis, R., Kapsos, S. & Kuhn, S. World employment and social outlook: trends 2020. ILO: Geneva Switzerland, 127. (2020).The World Bank. Employment in agriculture (% of total employment). (2025). https://data.worldbank.org/indicator/SL.AGR.EMPL.ZS?locations=NPWorld Bank. Unlocking Nepal’s Growth Potential: Nepal Country Economic Memorandum 2025. (2025b). Retrieved from https://www.worldbank.org/en/country/nepal/publication/unlocking-nepal-s-growth-potentialJaquet, S., Shrestha, G., Kohler, T. & Schwilch, G. The effects of migration on livelihoods, land management, and vulnerability to natural disasters in the Harpan watershed in Western Nepal. Mt. Res. Dev. 36 (4), 494–505. https://doi.org/10.1659/MRD-JOURNAL-D-16-00034.1 (2016).
    Google Scholar 
    Bhawana, K. C. & Race, D. Women’s approach to farming in the context of feminization of agriculture: A case study from the middle hills of Nepal. World Dev. Perspect. 20, 100260. https://doi.org/10.1016/j.wdp.2020.100260 (2020).
    Google Scholar 
    Adhikari, J. & Hobley, M. Everyone is leaving. Who will Sow our fields? The livelihood effects on women of male migration from Khotang and Udaypur districts, Nepal, to the Gulf countries and Malaysia. Himalaya 35 (1), 11–23 (2015).
    Google Scholar 
    National Statistics Office (NSO). National Population and Housing Census 2021: National Report (NSO, 2022).Maharjan, A., Bauer, S. & Knerr, B. Do rural women who stay behind benefit from male out-migration? A case study in the hills of Nepal. Gend. Technol. Dev. 16 (1), 95–123. https://doi.org/10.1177/097185241101600105 (2012).
    Google Scholar 
    Maharjan, A., Bauer, S. & Knerr, B. International migration, remittances and subsistence farming: evidence from Nepal. Int. Migration. 51, e249–e263. https://doi.org/10.1111/j.1468-2435.2012.00767.x (2013).
    Google Scholar 
    Upreti, B. R., Ghale, Y., Shivakoti, S. & Acharya, S. Feminization of agriculture in the Eastern hills of nepal: A study of women in cardamom and ginger farming. SAGE Open. 8 (4), 2158244018817124. https://doi.org/10.1177/2158244018817124 (2018).
    Google Scholar 
    Koirala, S. & Bashyal, S. The land left behind: a systematic review of transnational migration-induced change and its implication for rural sustainability in Nepal. Humanit. Social Sci. Commun. 12 (1), 1–12. https://doi.org/10.1057/s41599-024-04180-1 (2025).
    Google Scholar 
    Atreya, K. & Gartaula, H. Changing gender role declines maize yield, but remittances offset: findings from migrant households in the central Himalayas. Nepal. Outlook Agric. 51 (2), 247–259. https://doi.org/10.1177/00307270221097984 (2022).
    Google Scholar 
    Koirala, S. Empowering absence? Assessing the impact of transnational male Out-Migration on left behind wives. Social Sci. 12 (2), 80. https://doi.org/10.3390/socsci12020080 (2023).
    Google Scholar 
    Leder, S., Upadhyaya, R., van der Geest, K., Adhikari, Y. & Büttner, M. Rural out-migration and water governance: gender and social relations mediate and sustain irrigation systems in Nepal. World Dev. 177, 106544. https://doi.org/10.1016/j.worlddev.2024.106544 (2024).
    Google Scholar 
    Buisson, M. C., Clement, F. & Leder, S. Women’s empowerment and the will to change: evidence from Nepal. J. Rural Stud. 94, 128–139. https://doi.org/10.1016/j.jrurstud.2022.06.005 (2022).
    Google Scholar 
    Pandey, R. Farmers’ perception on agro-ecological implications of climate change in the Middle-Mountains of nepal: a case of Lumle village. Kaski Environ. Dev. Sustain. 21 (1), 221–247. https://doi.org/10.1007/s10668-017-0031-9 (2019).
    Google Scholar 
    Pandey, R. Male out-migration from the himalaya: implications in gender roles and household food (in) security in the Kaligandaki Basin, Nepal. Migration Dev. 10 (3), 313–341. https://doi.org/10.1080/21632324.2019.1634 (2021).
    Google Scholar 
    Kellogg, N. J., Andersen, A. B., Biraschi, R. & Puri, S. Beyond the (dis) empowerment binary: inevitability and the feminization of agriculture in Chitwan, Nepal. HIMALAYA-The J. Association Nepal. Himal. Stud. 40 (2), 108–118 (2021).
    Google Scholar 
    Yokying, P., Saksena, S. & Fox, J. Impacts of migration on time allocation of those who remain at home in rural Nepal. J. Int. Dev. 35 (7), 2067–2106. https://doi.org/10.1002/jid.3765 (2023).
    Google Scholar 
    Maharjan, A., Kochhar, I., Chitale, V. S., Hussain, A. & Gioli, G. Understanding rural outmigration and agricultural land use change in the Gandaki basin. Nepal. Appl. Geogr. 124, 102278. https://doi.org/10.1016/j.apgeog.2020.102278 (2020).
    Google Scholar 
    Maharjan, K. L. & Gonzalvo, C. M. Examining farmers’ willingness to learn environmental conservation agriculture: implications for women farmer empowerment in Bagmati Province. Nepal. Agric. 15 (7), 726. https://doi.org/10.3390/agriculture15070726 (2025).
    Google Scholar 
    Sugden, F. et al. Agrarian stress and climate change in the Eastern gangetic plains: gendered vulnerability in a stratified social formation. Glob. Environ. Change. 29, 258–269. https://doi.org/10.1016/j.gloenvcha.2014.10.008 (2014).
    Google Scholar 
    Gartaula, H. N., Niehof, A. & Visser, L. Feminisation of agriculture as an effect of male Out-migration: unexpected outcomes from Jhapa District, Eastern Nepal. Int. J. Interdisciplinary Social Sci. 5 (2). https://doi.org/10.18848/1833-1882/CGP/v05i02/51588 (2010).Kaspar, H. I Am the Household Head Now! Gender Aspects of Out-migration for Labour in Nepal (Nepal Institute of Development Studies, 2005).Kim, J. J., Stites, E., Webb, P., Constas, M. A. & Maxwell, D. The effects of male out-migration on household food security in rural Nepal. Food Secur. 11, 719–732. https://doi.org/10.1007/s12571-019-00919-w (2019).
    Google Scholar 
    Doss, C. R., Meinzen-Dick, R., Pereira, A. & Pradhan, R. Women’s empowerment, extended families and male migration in nepal: insights from mixed methods analysis. J. Rural Stud. 90, 13–25. https://doi.org/10.1016/j.jrurstud.2022.01.003 (2022).
    Google Scholar 
    Nadeem, M., Wang, Z., Shahbaz, P., Haq, S. U. & Boz, I. The nexus of women’s empowerment and sustainable agricultural practices in developing countries: a case of Pakistani women farmers. J. Environ. Planning Manage. 68 (4), 820–842. https://doi.org/10.1080/09640568.2023.2273198 (2025).
    Google Scholar 
    Shahbaz, P. et al. Adoption of climate smart agricultural practices through women involvement in decision making process: exploring the role of empowerment and innovativeness. Agriculture 12 (8), 1161. https://doi.org/10.3390/agriculture12081161 (2022).
    Google Scholar 
    Aryal, J. P. et al. Does women’s participation in agricultural technology adoption decisions affect the adoption of climate-smart agriculture? Insights from Indo‐Gangetic plains of India. Rev. Dev. Econ. 24 (3), 973–990. https://doi.org/10.1111/rode.12670 (2020).
    Google Scholar 
    Dar, M. H. et al. Gender focused training and knowledge enhances the adoption of climate resilient seeds. Technol. Soc. 63, 101388. https://doi.org/10.1016/j.techsoc.2020.101388 (2020).
    Google Scholar 
    Agarwal, T. et al. Gendered impacts of climate-smart agriculture on household food security and labor migration: insights from Bihar, India. Int. J. Clim. Change Strateg. Manag. 14 (1), 1–19. https://doi.org/10.1108/IJCCSM-01-2020-0004 (2022).
    Google Scholar 
    Glazebrook, T. & Opoku, E. Gender and sustainability: learning from women’s farming in Africa. Sustainability 12 (24), 10483. https://doi.org/10.3390/su122410483 (2020).
    Google Scholar 
    Theriault, V., Smale, M. & Haider, H. How does gender affect sustainable intensification of cereal production in the West African sahel? Evidence from Burkina Faso. World Dev. 92, 177–191. https://doi.org/10.1016/j.worlddev.2016.12.003 (2017).
    Google Scholar 
    DFID. Key sheets for sustainable development: overview. (1999).Natarajan, N., Newsham, A., Rigg, J. & Suhardiman, D. A sustainable livelihoods framework for the 21st century. World Dev. 155, 105898. https://doi.org/10.1016/j.worlddev.2022.105898 (2022).
    Google Scholar 
    UNDP. Guidance Note: Application of the Sustainable Livelihoods Framework in Development Projects (United Nations Development Programme, Regional Centre for Latin America and the Caribbean, 2017).Crenshaw, K. Mapping the margins: Intersectionality, identity politics, and violence against women of color. Stanford Law Rev. 43 (6), 1241–1299. https://doi.org/10.2307/1229039 (1991).
    Google Scholar 
    Collins, P. H. Black Feminist Thought, 2nd edition, New York: Routledge. (2000).Collins, P. H. Black Feminist Thought: Knowledge, consciousness, and the Politics of Empowerment (routledge, 2022).Kimberly, C. Demarginalizing the Intersection of Race and Sex: A Black Feminist Critique of Anti-Discrimination Doctrine, Feminist Theory and Anti‐Racist Politics. In The University of Chicago Legal Forum (Vol. 140, p. 139). (1989).Naess, L. O., Wangari-Muneri, E., Nightingale, A. J. & Mehta, L. Climate change and the operation of power: intersectionality, dispossession, and knowledge politics in pastoral communities. J. Peasant Stud. 1–26. https://doi.org/10.1080/03066150.2025.2451288 (2025).Tiwari, P. C. & Joshi, B. Gender processes in rural out-migration and socio-economic development in the himalaya. Migration Dev. 5 (2), 230–250. https://doi.org/10.1080/21632324.2015.1022970 (2016).
    Google Scholar 
    Islam, M. S., Sarker, M. F. H., Ehsan, S. M. A. & Sohel, M. S. Rethinking women empowerment in rural bangladesh: male out-migration, left-behind wives, and changing gender roles. Social Sci. Humanit. Open. 11, 101425. https://doi.org/10.1016/j.ssaho.2025.101425 (2025).
    Google Scholar 
    Pattnaik, I., Lahiri-Dutt, K., Lockie, S. & Pritchard, B. The feminization of agriculture or the feminization of agrarian distress? Tracking the trajectory of women in agriculture in India. J. Asia Pac. Econ. 23 (1), 138–155. https://doi.org/10.1080/13547860.2017.1394569 (2018).
    Google Scholar 
    Collins, A. Saying all the right things? Gendered discourse in climate-smart agriculture. J. Peasant Stud. 45 (1), 175–191. https://doi.org/10.1080/03066150.2017.1377187 (2017).
    Google Scholar 
    Mehta, D., Pandey, R., Gupta, A. K. & Juhola, S. Nature-based solutions in Hindu Kush himalayas: IUCN global standard based synthesis. Ecol. Ind. 154, 110875. https://doi.org/10.1016/j.ecolind.2023.110875 (2023).
    Google Scholar 
    Thapa, S. & Bhattarai, K. Farming Systems, Food Security and Contemporary Climate Issues in Nepal. In The Routledge International Handbook of Himalayan Environments, Development and Wellbeing (pp. 177–189). Routledge. (2025). https://doi.org/10.4324/9781003450894Sujakhu, N. M., Ranjitkar, S., Su, Y., He, J. & Xu, J. A gendered perspective on climate change adaptation strategies: a case study from Yunnan, China. Local Environ. 28 (1), 117–133. https://doi.org/10.1080/13549839.2022.2130883 (2023).
    Google Scholar 
    Raj, R., Ravula, P., Bhanjdeo, C. M. P., Sogani, A., Rao, N. & R., & Male migration and the transformation of gendered agriculture work: a comparative exploration of heterogeneity across selected Indian States. Gend. Place Cult. 1–27. https://doi.org/10.1080/0966369X.2025.2468178 (2025).Southard, E. M. & Randell, H. Climate change, agrarian distress, and the feminization of agriculture in South Asia. Rural Sociol. 87 (3), 873–900. https://doi.org/10.1111/ruso.12439 (2022).
    Google Scholar 
    Coxe, S., West, S. G. & Aiken, L. S. The analysis of count data: A gentle introduction to Poisson regression and its alternatives. J. Pers. Assess. 91 (2), 121–136. https://doi.org/10.1080/00223890802634175 (2009).
    Google Scholar 
    Hilbe, J. M. Poisson regression. In Modeling Count Data (35–73). Cambridge University Press. (2014).Kroll, C. N. & Song, P. Impact of multicollinearity on small sample hydrologic regression models. Water Resour. Res. 49 (6), 3756–3769. https://doi.org/10.1002/wrcr.20315 (2013).
    Google Scholar 
    García García, C., Gomez, S., García Pérez, J. & R., & A review of ridge parameter selection: minimization of the mean squared error vs. mitigation of multicollinearity. Commun. Statistics-Simulation Comput. 53 (8), 3686–3698. https://doi.org/10.1080/03610918.2022.2110594 (2024).
    Google Scholar 
    Long, J. S. & Freese, J. Regression Models for Categorical Dependent Variables Using Stata 3rd edn (Stata, 2014).Cameron, A. C., Trivedi, P. K. & Rev Microeconometrics using Stata (.) Stata. (2010).Meinzen-Dick, R., Quisumbing, A., Behrman, J., Biermayr-Jenzano, P., Wilde, V., Noordeloos,M., … Beintema, N. (2019). Engendering agricultural research. Food Security, 11(3),565–574. https://doi.org/10.1007/s12571-019-00917-6.Kandel, G. P., Bavorova, M., Ullah, A. & Pradhan, P. Food security and sustainability through adaptation to climate change: lessons learned from Nepal. Int. J. Disaster Risk Reduct. 101, 104279. https://doi.org/10.1016/j.ijdrr.2024.104279 (2024).
    Google Scholar 
    Kilic, T., Palacios-Lopez, A. & Goldstein, M. Caught in the productivity trap: A gendered perspective on the commercialization of agriculture in Africa. World Bank. Economic Rev. 36 (1), 75–104. https://doi.org/10.1093/wber/lhab006 (2022).
    Google Scholar 
    Nanda, M. & Ghosh, S. How does male out-migration impact the lives of left-behind women? Trade-off between feminization of agriculture and empowerment of farm women. Eval. Program Plan. 102603. https://doi.org/10.1016/j.evalprogplan.2025.102603 (2025).Fertő, I. & Bojnec, Š. Empowering women in sustainable agriculture. Sci. Rep. 14 (1), 7110. https://doi.org/10.1038/s41598-024-57933-y (2024).
    Google Scholar 
    Tourtelier, C., Gorman, M. & Tracy, S. Influence of gender on the development of sustainable agriculture in France. J. Rural Stud. 101, 103068. https://doi.org/10.1016/j.jrurstud.2023.103068 (2023).
    Google Scholar 
    Zhang, Y., Hu, N., Yao, L., Zhu, Y. & Ma, Y. The role of social network embeddedness and collective efficacy in encouraging farmers’ participation in water environmental management. J. Environ. Manage. 340, 117959. https://doi.org/10.1016/j.jenvman.2023.117959 (2023).
    Google Scholar 
    Beaman, L. & Dillon, A. Do female agricultural extension agents affect female farmers’ outcomes? Evidence from Ethiopia. Am. Economic Journal: Appl. Econ. 14 (1), 66–95. https://doi.org/10.1257/app.20200265 (2022).
    Google Scholar 
    Doss, C., Meinzen-Dick, R., Quisumbing, A. & Theis, S. Women in agriculture: four Myths. Global Food Secur. 16, 69–74. https://doi.org/10.1016/j.gfs.2017.10.001 (2018).
    Google Scholar 
    Ragasa, C., Aberman, N. L. & Alvarez-Mingote, C. Does providing agricultural and nutrition information to both men and women improve household food security? J. Agric. Econ. 70 (3), 940–963. https://doi.org/10.1111/1477-9552.12314 (2019).
    Google Scholar 
    Spangler, K. & Christie, M. E. Renegotiating gender roles and cultivation practices in the Nepali mid-hills: unpacking the feminization of agriculture. Agric. Hum. Values. 37, 415–432. https://doi.org/10.1007/s10460-019-09997-0 (2020).
    Google Scholar 
    Holmelin, N. B. Competing gender norms and social practice in Himalayan farm management. World Dev. 122, 85–95. https://doi.org/10.1016/j.worlddev.2019.05.018 (2019).
    Google Scholar 
    Fry, C. et al. Migrants as sustainability actors: contrasting nation, City and migrant discourses and actions. Glob. Environ. Change. 87, Article 102860. https://doi.org/10.1016/j.gloenvcha.2024.102860 (2024).Head, L., Klocker, N., Dun, O., Waitt, G., Goodall, H., Aguirre-Bielschowsky, I.,… Toole, S. (2021). Barriers to and enablers of sustainable practices: insights from ethnic minority migrants. Local Environment, 26(5), 595–614. https://doi.org/10.1080/13549839.2021.1904856.Batista, C., Han, D., Haushofer, J., Khanna, G., McKenzie, D., Mobarak, A. M., … Yang,D. (2025). Brain drain or brain gain? Effects of high-skilled international emigration on origin countries. Science, 388(6749), eadr8861. https://www.science.org/doi/full/10.1126/science.adr8861.Tambo, J. A. & Wünscher, T. Migration and conservation agriculture in Northern ghana: exploring the role of social networks and environmental perceptions. Food Secur. 9 (4), 685–698. https://doi.org/10.1007/s12571-017-0682-1 (2017).
    Google Scholar 
    De Brauw, A. et al. Biofortification, crop adoption and health information: impact pathways in Mozambique and Uganda. Am. J. Agric. Econ. 100 (3), 906–930. https://doi.org/10.1093/ajae/aay005 (2018).
    Google Scholar 
    Ram Mohan, R., Puskur, R., Chandrasekar, D. & Valera, H. G. A. Do gender dynamics in intra-household decision making shift with male migration? Evidence from rice-farming households in Eastern India. Gend. Technol. Dev. 27 (2), 157–183. https://doi.org/10.1080/09718524.2022.2140381 (2022).
    Google Scholar 
    Aryal, J. P., Thapa, G. & Simtowe, F. Mechanisation of small-scale farms in South asia: empirical evidence derived from farm households survey. Technol. Soc. 65, 101591. https://doi.org/10.1016/j.techsoc.2021.101591 (2021).
    Google Scholar 
    Khatri, N. D., Paudel, D., Bhusal, P., Ghimire, S. & Bhandari, B. Determinants of farmers’ decisions to adopt agroforestry practices: insights from the Mid-hills of Western Nepal. Agroforest Syst. 97 (5), 833–845. https://doi.org/10.1007/s10457-023-00830-6 (2023).
    Google Scholar 
    Bhatta, L. D., Shrestha, A., Neupane, N., Jodha, N. S. & Wu, N. Shifting dynamics of nature, society and agriculture in the Hindu Kush himalayas: perspectives for future mountain development. J. Mt. Sci. 16 (5), 1133–1149. https://doi.org/10.1007/s11629-018-5146-4 (2019).
    Google Scholar 
    Stark, O. & Bloom, D. E. The new economics of labor migration. Am. Econ. Rev. 75 (2), 173–178 (1985). https://www.jstor.org/stable/1805591
    Google Scholar 
    Yu, Z., Zhang, K., Wang, Z. & Liu, C. Immigration remittances, agricultural investment, and household wealth accumulation. Finance Res. Lett. 68, 105991. https://doi.org/10.1016/j.frl.2024.105991 (2024).
    Google Scholar 
    Mutai, N. C., Ibeh, L., Nguyen, M. C., Kiarie, J. W. & Ikamari, C. Sustainable economic development in kenya: influence of diaspora remittances, foreign direct investment and imports. Afr. J. Economic Manage. Stud. 16 (1), 61–78. https://doi.org/10.1108/AJEMS-01-2024-0059 (2025).
    Google Scholar 
    Wu, X., Chen, L., Ma, L., Cai, L. & Li, X. Return migration, rural household investment decision, and poverty alleviation: evidence from rural Guangdong, China. Growth Change. 54 (1), 304–325. https://doi.org/10.1111/grow.12656 (2023).
    Google Scholar 
    Cui, H., Wang, Y. & Zheng, L. Livelihood sustainability of rural households in response to external shocks, internal stressors and geographical disadvantages: empirical evidence from rural China. Environ. Dev. Sustain. 1–30. https://doi.org/10.1007/s10668-024-04666-7 (2024).Jijin, P. Remittances and labour force heterogeneity: can the disincentive effect vary? Int. Rev. Appl. Econ. 1–23. https://doi.org/10.1080/02692171.2024.2404947 (2024).Vemireddy, V. & Choudhary, A. A systematic review of labor-saving technologies: implications for women in agriculture. Global Food Secur. 29, 100541. https://doi.org/10.1016/j.gfs.2021.100541 (2021).
    Google Scholar 
    Arslan, A., McCarthy, N., Lipper, L., Asfaw, S. & Cattaneo, A. Adoption and intensity of adoption of conservation farming practices in Zambia. Agricultural Ecosyst. Environ. 254, 124–134. https://doi.org/10.1016/j.agee.2017.11.025 (2020).
    Google Scholar 
    Slavchevska, V., de la Campos, O., Brunelli, A. P., Doss, C. & C., & Beyond ownership: women’s and men’s land rights in sub-Saharan Africa. Food Policy. 102, 102065. https://doi.org/10.1016/j.foodpol.2021.102065 (2021).
    Google Scholar 
    Marty, E., Bullock, R., Cashmore, M., Crane, T. & Eriksen, S. Adapting to climate change among transitioning Maasai pastoralists in Southern kenya: an intersectional analysis of differentiated abilities to benefit from diversification processes. J. Peasant Stud. 50 (1), 136–161. https://doi.org/10.1080/03066150.2022.2121918 (2022).
    Google Scholar 
    Oyetunde-Usman, Z., Olagunju, K. O. & Ogunpaimo, O. R. Determinants of adoption of multiple sustainable agricultural practices among smallholder farmers in Nigeria. Int. Soil. Water Conserv. Res. 9 (2), 241–248. https://doi.org/10.1016/j.iswcr.2020.10.007 (2021).
    Google Scholar 
    Mgomezulu, W. R., Edriss, A. K., Machira, K. & Pangapanga-Phiri, I. Towards sustainability in the adoption of sustainable agricultural practices: implications on household poverty, food and nutrition security. Innov. Green. Dev. 2 (3), 100054. https://doi.org/10.1016/j.igd.2023.100054 (2023).
    Google Scholar 
    Adigun, G. T. Determinants of credit access among smallholder women farmers in Kwara State, Nigeria. Nigeria Agricultural J. 53 (2), 121–128 (2022).
    Google Scholar 
    Fares, M. H., Raza, S. & Ahmad, T. I. Complementarity in mixed farming systems enhances the smallholders income: evidence from Punjab, Pakistan. PloS One. 20 (4), e0319995. https://doi.org/10.1371/journal.pone.0319995 (2025).
    Google Scholar 
    Achmad, B., Sanudin, Siarudin, M., Widiyanto, A., Diniyati, D., Sudomo, A., Hani,A., Fauziyah, E., Suhaendah, E., Widyaningsih, T. S., Handayani, W., Maharani, D.,Suhartono, Palmolina, M., Swestiani, D., Budi Santoso Sulistiadi, H., Winara, A.,Nur, Y. H., Diana, M., … Ruswandi, A. (2022). Traditional Subsistence Farming of Smallholder Agroforestry Systems in Indonesia: A Review. Sustainability, 14(14), 8631. https://doi.org/10.3390/su14148631.Roy, D. et al. Impact of long term conservation agriculture on soil quality under cereal based systems of North West India. Geoderma 405, 115391. https://doi.org/10.1016/j.geoderma.2021.115391 (2022).
    Google Scholar 
    Chaudhuri, S. et al. Reflections on farmers’ social networks: a means for sustainable agricultural development? Environ. Dev. Sustain. 23, 2973–3008. https://doi.org/10.1007/s10668-020-00762-6 (2021).
    Google Scholar 
    Kiros, S. & Meshesha, G. B. Factors affecting farmers’ access to formal financial credit in Basona Worana District, North Showa Zone, Amhara regional State, Ethiopia. Cogent Econ. Finance. 10 (1). https://doi.org/10.1080/23322039.2022.2035043 (2022).Habib, N., Alauddin, M., Cramb, R. & Rankin, P. A differential analysis for men and women’s determinants of livelihood diversification in rural rain-fed region of pakistan: an ordered logit model (OLOGIT) approach. Social Sci. Humanit. Open. 5 (1), 100257. https://doi.org/10.1016/j.ssaho.2022.100257 (2022).
    Google Scholar 
    Download referencesFundingWe would like to express our gratitude to the Faculty of Tropical AgriSciences of the Czech University of Life Sciences in Prague for financially supporting this research (IGA 20223113). This project has also received funding from the European Union’s Horizon Europe research and innovation programme under grant agreement No 101084201 (ECO-READY). Moreover, we acknowledge funding from the European Research Council for the BeyondSDG project, the European Union (101077492). The sponsors had no role in the study design, data collection and analysis, decision to publish, or preparation of this manuscript.Author informationAuthors and AffiliationsFaculty of Agrobiology, Food, and Natural Resources, Czech University of Life Sciences Prague, Kamycka 129, Prague, Suchdol, 16500, Czech RepublicGiri Prasad Kandel & Ioannis ManikasFaculty of Tropical AgriSciences, Czech University of Life Sciences Prague, Kamycka 129, Prague, Suchdol, 16500, Czech RepublicMustapha Yakubu Madaki, Tereza Pilarova, Ayat Ullah & Miroslava BavorovaEnergy and Sustainability Research Institute Groningen (ESRIG), University of Groningen, Groningen, 9747 AG, NetherlandsPrajal PradhanPotsdam Institute for Climate Impact Research (PIK), Member of the Leibniz Association, P.O. Box 60 12 03, Potsdam, D-14412, GermanyPrajal PradhanAuthorsGiri Prasad KandelView author publicationsSearch author on:PubMed Google ScholarMustapha Yakubu MadakiView author publicationsSearch author on:PubMed Google ScholarTereza PilarovaView author publicationsSearch author on:PubMed Google ScholarAyat UllahView author publicationsSearch author on:PubMed Google ScholarPrajal PradhanView author publicationsSearch author on:PubMed Google ScholarIoannis ManikasView author publicationsSearch author on:PubMed Google ScholarMiroslava BavorovaView author publicationsSearch author on:PubMed Google ScholarContributionsGiri Prasad Kandel: Conceptualization; Methodology; Data curation; Formal analysis; Writing – original draft; Project administration.Mustapha Yakubu Madaki: Methodology; Validation; Writing – review & editing.Tereza Pilarova: Methodology, Validation, Writing – review & editing.Ayat Ullah: Formal analysis; Visualization; Writing – review & editing.Prajal Pradhan: Supervision; Writing – review & editing.Ioannis Manikas: Supervision; Funding acquisition, Writing – review & editing.Miroslava Bavorova: Conceptualization; Supervision; Funding acquisition; Writing – review & editing.Corresponding authorCorrespondence to
    Giri Prasad Kandel.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-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 articleKandel, G.P., Madaki, M.Y., Pilarova, T. et al. Gender dynamics and remittances in the adoption of sustainable agricultural practices in Nepal.
    Sci Rep (2025). https://doi.org/10.1038/s41598-025-31848-8Download citationReceived: 29 August 2025Accepted: 05 December 2025Published: 15 December 2025DOI: https://doi.org/10.1038/s41598-025-31848-8Share 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
    KeywordsFeminization of agricultureGender dynamicsMale out-migrationClimate changeRemittancesSustainable agricultural practices More

  • in

    Giving a voice to animals: Laos’s national herpetologist on her day-to-day

    “Because I am one of very few specialists in Laos, some people have started to describe me as the country’s first national herpetologist.This is accurate to my knowledge; I’m not aware of any other Laotians who were working in the field here before 2007, when I started my master’s degree. Conservation studies were instead almost always done by visiting foreign scientists. My PhD challenged that trend, however: I focused on the diversity of Tylototriton salamanders, which led to the discovery of four species.My day-to-day responsibilities at the National University of Laos in Vientiane include giving lectures, conducting research, leading studies, training future biologists and contributing to conservation initiatives. My own research focuses on amphibian and reptile diversity, with an emphasis on conserving threatened species in the country.In this photograph, I’m surveying a population of Fejervarya limnocharis frogs along a stream outside Vientiane, for a field survey. I’m recording the number of adults, tadpoles, eggs and frog calls, as well as tracking other ecological factors.

    Enjoying our latest content?
    Log in or create an account to continue

    Access the most recent journalism from Nature’s award-winning team
    Explore the latest features & opinion covering groundbreaking research

    Access through your institution

    or

    Sign in or create an account

    Continue with Google

    Continue with ORCiD More

  • in

    Compounding preconditions of wildfires vary in time and space within Europe

    AbstractFavorable wildfire conditions are increasing in frequency and severity across Europe. Understanding how wildfire drivers vary in space and time is crucial for mitigating wildfire risk under current and future climate conditions. Here, we analyze the hydro-meteorological and land-surface drivers of wildfires from 2001 to 2020 across eight European climate regions and their mountain ranges. Our findings reveal that drought conditions and vapor pressure deficit are the dominant drivers of wildfire activity. These drivers vary by season and region: in Southern and Central European regions, persistent warm and dry conditions in preceding seasons favor summer wildfires, while fall wildfires are influenced by fuel build up in spring that loses moisture during dry and hot summer weather. In Northern Europe, these dynamics occur on sub-monthly timescales. Our results illustrate the critical role of compounding wildfire drivers and emphasize the need for targeted mitigation strategies, especially in the light of climate change.

    Similar content being viewed by others

    Substantial increase in wildfire danger conditions under anthropogenic climate change in Southwest France

    Article
    Open access
    23 July 2025

    Analysing historical events and current management strategies of wildfires in Norway

    Article
    Open access
    10 July 2025

    Future enhanced threshold effects of wildfire drivers could increase burned areas in northern mid- and high latitudes

    Article
    Open access
    21 March 2025

    IntroductionThe wildfire seasons of 2023 and 2022 were ranked among the five worst wildfire seasons of the European continent since the beginning of the century1,2. Even though burned area shows a long-term decline in the Mediterranean region3,4 and at the global scale5, burned area increases in mid- and high-latitude climate regions4. In these regions, increasing wildfire activity is especially concerning, because of abundant fuel availability6, and lower adaptation and resilience of plants to fires7.In recent years, Central and Northern Europe have experienced severe wildfires, promoted by intense drought conditions that led to high dry fuel availability in a usually moisture-limited fire regime8, e.g., in Northern Europe during the drought in 20189,10 and in Central Europe during the drought of 20222,11. Under climate change, warmer and drier climate zones are expected to expand towards northern European latitudes12,13, enhancing fire weather and dry fuel availability, which consequently increases the risk of large wildfires in these regions14,15,16. As wildfire-favoring conditions intensify in historically less wildfire-prone areas, and wildfire seasons increase in length in already wildfire-prone regions of Europe17, a more detailed understanding of the individual and compounding wildfire preconditions, and their temporal persistence across seasons and climate regions of Europe is needed.Wildfires can be understood as compound events that result from spatially and temporally compounding hydro-meteorological and land-surface conditions that favor wildfire ignition and spread18. These conditions include both meteorological conditions, such as high temperatures, strong vapor pressure deficits (VPDs), and high wind speeds4, as well as land-surface conditions such as drought conditions resulting from precipitation and soil moisture deficits, which modulate fuel loads and availability4,8. Although there is a strong link between meteorological fire weather described by fire weather indices, such as the Canadian Forest Fire Weather Index (FWI)19, and wildfire occurrence20,21, fuel loads and availability are not explicitly represented in these index systems22. Still, the FWI accounts for fuel moisture in its fuel moisture codes, which empirically capture drought conditions in different fuel layers for up to 52 days19. We hypothesize that wildfire preconditions vary by season and region, and that understanding these characteristics requires analyzing individual hydro-meteorological and land-surface variables, such as gross primary productivity, soil moisture, and snow depth, rather than solely relying on compound indices like the FWI. Therefore, we here examine each hydro-meteorological and land-surface variable separately, without assuming any a priori compounding relationship between the individual wildfire drivers.The amount and timing of biomass accumulation are crucial for wildfire activity as gross primary productivity in wet periods modulates fuel availability for the following dry periods, e.g., the growth of Mediterranean shrubs in wet periods in Mediterranean climates8,23. In regions where fire activity is moisture-limited, prolonged dry periods decrease fuel moisture and the risk of large wildfires24,25. The processes of fuel accumulation and drying act—depending on the ecosystem—on seasonal to annual timescales23,26. In contrast, meteorologically distinct fire weather is characterized by hot, dry, and windy atmospheric conditions that occur on daily to weekly timescales and are influenced by large-scale weather patterns over Europe27,28.Fire weather, season length, and vegetation dynamics vary across environmental zones29, within which climate variability modulates annual burned area fluctuations30: In Mediterranean and more arid climates, pronounced antecedent wet periods can promote above-average biomass productivity, leading to higher fuel loads in subsequent dry periods that can result in extreme wildfire activity30,31,32,33, e.g., as observed in Portugal in 201726. In contrast, in temperate and boreal climate zones, where wildfire occurrence is moisture-limited, sub-seasonal drought conditions increase dry fuel availability, which in turn promotes higher wildfire activity than average climate conditions30,34, e.g., as in Sweden in 201835. In mountain regions, such as the southern Alps, prolonged dry periods and snow-scarce conditions in winter, in combination with strong katabatic winds, i.e., Foehn winds, promote snow-free fuel layers resulting in fast-spreading winter wildfires24,36. These examples highlight that inter-annual variability affects wildfire preconditions differently in each region and season. Therefore, a detailed analysis of climate–vegetation–wildfire interactions is needed to understand region- and season-specific wildfire danger.Until now, wildfires in Europe have been mostly studied on national and regional levels, with a focus on the highly wildfire-affected areas in Southern Europe. Recent severe wildfire seasons that coincided with strong drought conditions in Central and Northern Europe suggest that the “switch”8 for flammable conditions in moisture-limited regions is turned on. Given that wildfires and their severity are not exclusively modulated by short-term weather conditions, but also by antecedent wet-dry transitions, wildfire occurrences and their preconditions need to be studied from a holistic perspective. Such a holistic perspective should include land-surface parameters and hydro-meteorological drivers that go beyond traditional fire weather indices and consider daily, seasonal, and annual timescales. In this study, we highlight the importance of individual wildfire drivers for the occurrence of wildfire events in different European regions by using random forest models and assess when preconditions start becoming anomalous before a wildfire event. We analyze the three-month Standardized Precipitation Evaporation Index (SPEI-3M), soil moisture, and snow depth deficit as long-term wildfire drivers and temperature, VPD, and wind speed as short-term wildfire drivers. In addition, we consider gross primary productivity (GPP) (accumulated over 1 month) as a proxy for fuel drying, which is influenced by both, short- and long-term drivers. By analyzing all of these potential wildfire drivers individually, we are able to disentangle the compounding preconditions that lead to wildfire occurrence in Europe. To address the challenge of wildfire drivers acting on different timescales, we propose the concept of “time of precondition emergence” (ToPE). This concept will allow us to better understand the relative importance of and interplay between long- and short-term wildfire drivers, as well as the temporal onset of wildfire preconditions. Such understanding is highly relevant for early warning and wildfire prevention, especially under changing climate conditions.Here, we disentangle the regional and temporal characteristics of hydro-meteorological and land-surface drivers of wildfires in Europe. We divide Europe into eight different climate regions, namely the PRUDENCE regions37, and account for altitude-related climate gradients within these regions by distinguishing between high-mountain and non-mountain regions38. We study the regional and seasonal characteristics of wildfires and their drivers by (i) describing the seasonal and regional characteristics of wildfire occurrence in Europe, (ii) comparing the hydro-meteorological and land-surface conditions of days when wildfires start compared with days when no wildfires are burning (non-wildfire days), (iii) understanding to which degree these conditions differ between high-mountain and non-mountain regions, (iv) investigating regional differences in the most important drivers of wildfires, and (v) quantifying the length of the time period during which hydro-meteorological and land-surface conditions have been anomalous prior to wildfire events.Results and discussionSeasonal and regional characteristics of wildfiresWe study the spatial and temporal patterns of wildfires across eight regions [i.e., PRUDENCE regions37] and their respective high mountain ranges [i.e., high and scattered high mountains in ref. 38] by leveraging the ESA FireCCI version 5.1 (FireCCI51) burned area pixel product39 between 2001 and 2020. We assess the spatial distribution of wildfires by the cumulative burned area over the 20-year study period and describe the seasonality in burned area and number of wildfire events as each season’s fraction of the regional and mountain range types of the 20-year total.Wildfire occurrence and peak seasonality in Europe vary greatly across the eight examined regions and their mountain ranges (Fig. 1). Wildfires are most common in summer in Southern Europe (Iberian Peninsula and Mediterranean), France, the mountain regions of Eastern Europe and Scandinavia, and in the non-mountain parts of the Alps. Regions characterized by summer wildfires experience wildfires in fall as well, though the seasonality is less pronounced than in summer. In the fall, mountain regions are more strongly affected by wildfires than non-mountain regions, whereas in summer, non-mountain regions are more strongly affected. In the British Isles, the non-mountain regions of Scandinavia, Mid-Europe, and Eastern Europe, wildfire seasonality peaks in spring.Fig. 1: Spatial and temporal distribution of wildfires in Europe.a Cumulative burned area (2001–2020) per CERRA grid-cell (5.5 km) for Europe (BI British Isles, FR France, IP Iberian Peninsula, SC Scandinavia, ME Mid-Europe, AL Alps, MD Mediterranean, EA Eastern Europe). b Relative burned area (hatched) and number of wildfire events (dotted) of the annual mean sum derived for each region and its respective mountain (green) and non-mountain regions (yellow).Full size imageIn terms of total burned area over the 20-year observation period (i.e., 2001–2020), regions in Southern Europe are the most strongly affected by wildfires, with large burned area clusters in the western Iberian Peninsula, southern Italy, and the Balkan countries (see Fig. 1). The second region most strongly affected by wildfires is Eastern Europe, where wildfires occur along the Carpathian mountain range, but are most dominant on the border between Ukraine and Belarus, as well as the Russian exclave Kaliningrad. In Scandinavia, wildfires occur mostly in the non-mountain regions of Sweden and the Baltic states (i.e., Estonia, Lithuania, and Latvia), while mountain regions experience fewer wildfires. In contrast, the mountain ranges of the British Isles, including the Scottish Highlands, show frequent wildfire occurrence. Mid-Europe, France, and the Alps are the regions with the fewest observed wildfires, but each of these regions has at least one wildfire hotspot, i.e., the Atlantic Coast and Massif Central in France, the southern mountain ranges in the Alps, and the northern parts of Mid-Europe.Hydro-meteorological and land-surface conditions of wildfiresWe analyze hydro-meteorological and land-surface conditions of wildfire events in terms of seasonal and absolute anomalies in the eight PRUDENCE regions, subdivided into their mountain and non-mountain areas. We describe the conditions of wildfire events on their first day of detection, i.e., wildfire start days, and compare these conditions to days, when no wildfires are detected, i.e., non-wildfire days, using our event detection algorithm (see “FireCCI51 and derived wildfire events” under the “Methods” section). We exclude conditions on days that are not wildfire start days to avoid oversampling the same event in case it burns for multiple days. For each wildfire start day and non-wildfire day, we sample short-term drivers, i.e., maximum temperature, vapor pressure deficit (VPD), 10-m surface wind speed, and long-term drivers, i.e., 3-month Standardized Precipitation Evapotranspiration Index (SPEI-3M), soil moisture, and snow depth deficit, as well as gross primary productivity accumulated over 32 days (GPP-1M) as a proxy for fuel drying. We apply a Mann–Whitney U-test to test whether there is a significant difference between wildfire start and non-wildfire days, and between mountain and non-mountain regions.Wildfires occur in all seasons and regions under warmer and drier conditions than the seasonal and regional mean (Fig. 2). These seasonal hot and dry conditions likely affect GPP-1M, which reflects the availability of dry fuels, and hence affects region-specific wildfire activity (see Fig. 2). We find that in spring and summer, GPP-1M is less abundant on wildfire start days than on non-wildfire days in all regions except Scandinavia and the British Isles, where GPP-1M is more abundant on wildfire start days in comparison to non-wildfire days. In fall and winter, GPP-1M is more abundant in all regions on wildfire start days than on non-wildfire days. Seasonal temperature anomalies are less strongly pronounced than SPEI-3M, soil moisture and VPD anomalies (see Fig. 2). In summer, wildfire drivers on wildfire start days are less strongly pronounced in Southern Europe in comparison to other regions, which suggests that wildfire-favoring conditions might be met on non-wildfire days too (see Supplementary Material Fig. S2), or wildfire activity is related to other factors, which are not considered in this study. Seasonal wind speed is higher on wildfire start days in comparison to non-wildfire days in Southern Europe (Iberian Peninsula and Mediterranean) in summer, in fall in the British Isles and Alps, and in winter in the Alps. In contrast, wind speed in the remaining regions and seasons is lower on wildfire start days in comparison to non-wildfire days. We observe a large variability in wind speed on wildfire start days (see Supplementary Material Fig. S2) and assume that the 5.5 km spatial resolution limitedly resolves the 10-m daily surface maximum wind speed during initial wildfire ignition. Seasonal snow depth deficits are pronounced (i.e., ≥0.25 SD) on wildfire start days in winter in Northern and Eastern Europe and in France, the Alps, and Mid-Europe. In spring, the seasonal snow depth deficits on wildfire days are less strongly pronounced than in winter, and are only larger than +0.25 SD in the Alps and Scandinavia. While wildfires under these snow depth deficits most likely occur in snow-free periods, such snow-free periods are atypical in these regions and seasons, as shown by the comparison to non-wildfire days.Fig. 2: Standardized anomalies of hydro-meteorological and land-surface drivers on wildfire start days in comparison to non-wildfire days.We show 2-m maximum temperature (Tmax), 2-m vapor pressure deficit (VPD), 10-m surface wind speed (WindS), 32-day accumulated (GPP-1M), 3-month SPEI (SPEI-3M), soil moisture (SoilM), and snow depth (SnowD) on the start day of wildfires in comparison to days without wildfires for each region and season over the entire subregion (“Over all” panel; left), mountain regions (“Mountains panel”; middle left) and non-mountain regions (“Non-mountains”; middle right). Colored values are based on the seasonal median difference between all wildfire start events and all non-wildfire events. The right panel (“On wildfire start days”) shows the difference between mountain regions and non-mountain regions on wildfire start days. Hatching indicates that the difference between wildfire start and non-wildfire days (first three panels from left) and mountain vs non-mountain regions (right panel) is not significant (Mann–Whitney U-test, p-value ≤ 0.05). Gray panels indicate that no events are observed in this region and season. To have a uniform sign across all variables, we flip the sign for deficit variables and mark these with an asterisk (*), i.e., SPEI-3M soil moisture and snow depth.Full size imageAlthough the standardized seasonal anomalies of the examined wildfire drivers are distinct, looking at these drivers from an absolute perspective provides further insights: Absolute temperature and VPD in summer are high in Southern Europe on wildfire start days (see Fig. S2). However, we find that these variables are, from a seasonal perspective, not strongly different from those on non-wildfire days (see Fig. 2). Strong seasonal soil moisture deficits on wildfire start days reflect an absolute deficit of more than -0.5 SD in spring, summer, and fall. In contrast, seasonal soil moisture anomalies in winter originate from an absolute soil moisture surplus, which is still lower than on non-wildfire days (see Fig. S2). The SPEI-3M is less season-dependent, and the identified seasonal anomalies are of the same magnitude as the absolute anomalies (see Fig. S2).Different wildfire conditions in mountain vs non-mountain regionsWe find that most drivers of wildfires are more pronounced in mountain regions than in non-mountain regions (see Fig. 2), which indicates that wildfires in mountain regions start under seasonally stronger anomalies than wildfires in non-mountain regions. However, these differences are significant only in a few regions and seasons: On wildfire days we find that wind speed, snow depth deficit (e.g., Northern Europe in spring and the Alps in fall and winter) and SPEI-3M (e.g., Eastern Europe in summer) are significantly enhanced in mountain regions, when compared to non-mountain regions (p-value ≤ 0.05) (see Fig. 2, “On wildfire start days” panel). On wildfire start days, maximum temperature (e.g., British Isles and Alps in summer), GPP-1M (e.g., in Mid-Europe and Eastern Europe in spring and in Eastern and Southern Europe in winter), and SPEI-3M (e.g., in Scandinavia in summer and fall) show stronger anomalies for wildfire-favoring conditions in non-mountain than in mountain regions. These anomalies reflect the elevational and seasonal dependencies of wildfire drivers in the respective regions and seasons.Most important wildfire drivers by region and seasonWe identify the most important drivers of wildfire events on their start day for each region and season by utilizing the variable importance metric of random forest models that predict binary wildfire occurrence (see details in the chapter “Variable importance of wildfire drivers” under the “Methods” section). The overall accuracy of the confusion matrix of the random forest models, which were trained on a balanced wildfire start day vs non-wildfire day dataset, ranges between 0.76 and 0.96 (overall accuracy, see Supplementary Materials Fig. S3) and provides a good baseline to interpret variable importance derived from the Gini-Impurity index.The overall dominant drivers of wildfires are SPEI-3M, soil moisture deficit, and VPD. However, the relative importance of these wildfire drivers varies regionally and seasonally as they interact with other wildfire drivers, such as maximum temperature and GPP-1M. In spring and summer, the most important drivers of wildfire events are GPP-1M and VPD, while SPEI-3M and VPD are crucial in fall and winter (see Fig. 3). In spring, GPP-1M is the most important wildfire driver in all regions, except in the British Isles, France (ranked 2nd), and the Alps (ranked 3rd). In spring, the importance of GPP-1M as a wildfire driver reflects a seasonal deficit of GPP-1M in all regions (see Fig. 2). VPD and soil moisture deficit are the second and third most important variables in spring, as they are ranked among the three most important drivers in all regions. In summer, the most important drivers differ between the northwestern (i.e., France, Mid-Europe, British Isles, and Scandinavia) and southeastern (i.e., Iberian Peninsula, Alps, Mediterranean, and Eastern Europe) regions: in the southeastern regions, seasonal GPP-1M deficit is still the most important variable together with SPEI-3M and VPD. In the northwestern regions, VPD, maximum temperature, and a long-term drought variable (either SPEI-3M or soil moisture deficit), represent the most important wildfire drivers (see Fig. 3). In fall, long-term drought conditions, as described by the SPEI-3M and soil moisture deficit, along with VPD, are the most important variables for wildfire occurrence. In Southern Europe and Mid-Europe, elevated temperatures play an important role, too, while in the remaining regions, temperature is not ranked among the three most important variables. In winter, wildfires are most importantly influenced by VPD. Besides high VPD, long-term drought conditions reflected in the SPEI-3M are ranked among the three most important drivers of wildfires in the Iberian Peninsula, the British Isles, France, and the Alps. In addition, GPP-1M surpluses become crucial for wildfire occurrence in Southern Europe, France, and the British Isles. Positive temperature anomalies play a critical role in winter wildfire occurrence in Scandinavia, Mid-Europe, Eastern Europe, and the Mediterranean.Fig. 3: Drivers of wildfires ranked by random forest feature importance.a Three most important variables on the start day of wildfire events based on the Gini impurity index derived from seasonal and regional random forest models. b Variable importance by region and season for all drivers used in the random forest models.Full size imageTime of precondition emergenceWe investigate how long individual wildfire drivers have been anomalous before the start of wildfires. Specifically, we average the absolute anomalies of each driver over incrementally increasing time steps, to identify the time window during which conditions on days when wildfires start differ notably from conditions on non-wildfire days. The time steps considered range from the day of wildfire start, i.e., day 0, up to one year before the day of wildfire start, i.e., 360 days. We call the time step when wildfire start day anomalies emerge from 1 standard deviation (SD) of the difference between wildfire start day and non-wildfire day anomalies the “time of precondition emergence” (ToPE). This 1 SD of the difference is mapped onto the non-wildfire conditions to account for natural variability (see Fig. 4a). For example, in summer in the Mediterranean region, maximum temperature and VPD and SPEI-3M anomalies emerge 270 and 240 days before wildfire events start from non-wildfire conditions, respectively, while GPP-1M deficits emerge 60 days before wildfire events start (see Fig. 4b). The long ToPE times for maximum temperature, VPD and SPEI-3M, imply that preceding months have been anomalously warm, atmospherically dry, and precipitation-scarce. In summary, these findings show how the preconditions of wildfires develop over time. The example of Mediterranean summer shows how preceding warm and dry spring and winter months lead to the availability of dry fuels, i.e., a GPP-1M deficit exceeding 1 SD 60 days prior to wildfire start days (see Fig. 4b).Fig. 4: Time of precondition emergence (ToPE) for wildfire drivers: framework and results.a Conceptual framework developed to determine ToPE, illustrating how emergence levels of all driving variables are derived for a threshold of 1 and 2 SDs of the difference between wildfire and non-wildfire conditions. b Example of how ToPE is derived for different variables for the Mediterranean region in summer (JJA). If ToPE does not occur at the 1 or 2 SD threshold level, the dot is plotted at the bottom-right of the individual subplot (e.g., 2 SD for GPP-1M). c ToPE for all variables exceeding the threshold of 1 SD, which is derived from the difference between wildfire and non-wildfire values for the respective variable (color), season (panel), and region (y-axis). Variables with an asterisk (*) in the legend are deficit variables, hatched GPP-1M values represent deficits, and solid-filled GPP-1M values represent surpluses. Faded variables are not significantly different (p-value ≤ 0.05) on wildfire start days compared to non-wildfire days according to the Mann–Whitney U-test.Full size imageBy looking at all ToPE times for each variable, season and region, we find strong seasonal and regional patterns, with generally shorter ToPE times in spring and winter, and generally longer ToPE times in summer and fall (see Fig. 4c): In spring in Southern (Iberian Peninsula and Mediterranean) and Eastern Europe, SPEI-3M emerges from non-wildfire conditions more than 270 days before wildfire events occur. These drought conditions likely lead to a GPP-1M deficit that emerges 90 days before wildfire events start. VPD and snow-depth deficit emerge on a sub-monthly scale (ToPE ≤ 30 days) in these regions. In Northern Europe (Scandinavia and British Isles) and France, wildfire preconditions emerge up to 14 days prior to wildfire events (see Fig. 4c). In summer, we observe persistent wildfire preconditions longer than 90 days in all regions, except Mid-Europe and the British Isles. These persistent preconditions of high maximum temperatures and VPD, alongside pronounced soil moisture deficits and low SPEI-3M, likely lead to GPP-1M deficits, that develop 60 to 30 days prior to wildfires. ToPE times in fall are shorter in Southern Europe and France, longer in Eastern Europe, the Alps, and Mid-Europe, and equally long in Northern Europe in comparison to the summer season. We find that SPEI-3M is anomalously low in Mid-Europe, the Alps, and Eastern Europe a year before wildfire starts in fall, and likely leads to soil moisture deficits that emerge on shorter timescales than temperature and VPD. GPP-1M surpluses for fall wildfires emerge from non-wildfire conditions in the previous vegetation period, i.e., ToPE ≥ 90 days in France, the Iberian Peninsula, and Scandinavia. We show that for these regions, wildfires in fall have relatively short ToPE times for VPD, temperature, and soil moisture compared to ToPE for GPP-1M surpluses, leading to dry fuel availability. ToPE times ≥ 180 days for snow depth deficits in the Alps and the Mediterranean region, alongside of persistent (ToPE times ≥ 90 days) SPEI-3M conditions show that fall wildfires can be influenced by the previous season’s snow cover (see Fig. 4c). For winter wildfires, snow depth deficits develop on same-season timescales (ToPE ≤ 90 days) in the British Isles, Scandinavia, the Alps, France, Eastern Europe, and the Iberian Peninsula. In these regions, SPEI-3M is anomalously low for longer periods than snow depth deficits, and VPD and temperature are anomalously high on shorter timescales than snow depth deficits. We find that GPP-1M surpluses in winter are anomalously high from the previous active vegetation period, with ToPE times ≥ 240 days in France and Southern Europe. These findings illustrate how long-term drought conditions in combination with anomalously high VPD can promote wildfire danger in winter.Wildfires result from compounding driversOur results highlight how land-atmosphere interactions affect wildfire occurrence by illustrating how hydro-meteorological drivers and drought conditions affect fuel availability, i.e., available biomass to burn, approximated from GPP. We find that pronounced dry conditions, resulting from short-term atmospheric dryness (i.e., VPD) and long-term drought (i.e., SPEI-3M or soil moisture deficit), are the most important drivers of wildfire occurrence in Europe, whereas temperature and wind speed play a secondary role (see Fig. 2). We find that the seasonal anomalies for all variables vary mostly in the range of 0.25 and 1 SD and single variable outliers larger than 1 SD occur mostly in summer and fall. These findings highlight that the co-occurrence of multiple moderately pronounced drivers is more relevant for wildfire activity than strong anomalies in one variable alone, which is in line with findings for other regions of the world, such as California40.In Europe, the development of cascading wildfire preconditions starts with biomass drying (i.e., GPP-1M deficit) as biomass is sufficiently available even in the driest climate regions of Europe, i.e., Southern Europe8,23. The availability of dry fuel results from interactions between GPP-1M, temperature, and dry conditions, i.e., soil moisture deficits, low SPEI-3M, and VPDs (see Fig. 4). Wildfire occurrence is restricted by fuel availability in dry, or “fuel-limited”, regions such as the driest parts of Southern Europe, and is “moisture-limited” in regions with high GPP but sufficient moisture, such as Northern and Central Europe8. Our results of these interactions represent relative changes in the regional and seasonal live fuel moisture levels rather than the full region-specific fuel load and structure as in other studies that use a sophisticated fuel model [i.e., refs. 31,41,42]. Still, we illustrate how hydro-meteorological conditions influence the flammability of biomass (see Fig. 2): in spring and summer, GPP-1M is lower on wildfire start days than under non-wildfire conditions, indicating that VPD and soil moisture deficits likely lead to a decrease in GPP-1M and promote its drying relative to non-wildfire conditions. However, in Northern Europe (Scandinavia and the British Isles), high VPD, low SPEI-3M, and soil moisture deficits are associated with slightly increased (≤ 0.25 SD) GPP-1M on days when wildfires occur in summer. This suggests that GPP-1M decreases only when soil moisture is below a certain absolute threshold43. With respect to wildfire occurrence, our findings imply that GPP-1M deficits in spring and summer represent decreases in fuel moisture in Southern and Central Europe, while GPP-1M surpluses in Northern Europe under drought conditions44 become rapidly available as dry fuel under high VPD and high temperatures (see Figs. 2 and 4).Our results for the ToPE highlight memory effects of all driving variables from daily to annual timescales prior to wildfire events and illustrate the temporal development of compounding preconditions for wildfires in different seasons and regions (see Fig. 4). While regional differences are present, we identify distinct propagation patterns of wildfire preconditions in different seasons: In spring, ToPE times are short, suggesting that wildfire preconditions develop rapidly in this season. In Southern and Eastern Europe, preceding dry winter and fall conditions (SPEI-3M) additionally favor wildfires in spring (see Fig. 4c). In summer, ToPE times for maximum temperature, VPD, SPEI-3M, and soil moisture deficit are longer than 90 days in Southern Europe, Eastern Europe and the Alps showing that wildfire preconditions manifest early in the fire season, while wildfire preconditions develop on shorter timescales in the remaining regions (see Fig. 4c). For wildfires in fall, we demonstrate that fuel (i.e., GPP-1M surplus) is accumulated in spring (Mediterranean, Iberian Peninsula and France) and summer (Scandinavia), which likely becomes available as dry fuel under high temperatures and enhanced VPD, which start developing up to 30 days before wildfires start. For these regions, ToPE patterns are similar in the winter, however, ToPE is comparatively longer for GPP-1M and shorter for temperature, VPD, and soil moisture. In winter in Northern Europe (British Isles, Scandinavia), Eastern Europe, and the Alps, long-term drought conditions (SPEI-3M) lead to snow depth deficits that promote wildfires under high VPD. ToPE returns the longest time step prior to an extreme event for which a specific precondition variable starts to emerge from non-extreme event conditions. This time step does not necessarily represent the peak of the precondition anomaly, rather, it indicates its earliest onset, and, therefore, the start of the cascade of anomalous preconditions for wildfires. Our ToPE analysis provides insights into how quickly dangerous wildfire preconditions develop in different regions and seasons, which can inform wildfire forecasting and management.The role of wind and snow in wildfire occurrenceAccording to our analysis, the most important drivers of wildfire occurrence are long-term drought conditions and VPD throughout all seasons, while wind speed and snow depth are only tangentially important in most regions (see Figs. 3 and 2). For most regions and seasons, we find that wind speed is significantly lower on wildfire start than on non-wildfire days, except in Southern Europe (Iberian Peninsula and Mediterranean) in summer, the British Isles and the Alps in fall, and the Alps in winter (see Fig. 2 panel “Overall”). In the mountain regions of the Alps, wind speed is particularly relevant, because wind-driven wildfires in winter co-occurring with snow depth deficits explain the majority of the total annual burned area in this region24. Nevertheless, wind speed is a driver of wildfire size in other regions and seasons as well45, especially in regions with local wind systems, such as the “Mistral” in France46 or the “Melteni” in Greece4,47. Our results show a high variability of wind speed on wildfire start days (see Supplementary Material Fig. S2), suggesting that daily 10-m maximum surface wind speed in the 5.5 km CERRA grid is not sufficient to resolve the high sub-daily and spatial variability of wind speed on wildfire start days, and that wind speed might be more relevant for wildfire spread than ignition.Though the majority of wildfires occur in summer and fall across Europe, wildfires in spring and winter represent a substantial fraction of all wildfire events in Scandinavia, British Isles, Eastern Europe, France and the Alps (see Fig. 1b). Winter and spring wildfires in these regions are accompanied by snow depth deficits, high VPD and soil moisture deficits (see Fig. 2). In these snow-dominated regions, snow depth deficits likely lead to snow free periods, that create potential for wildfire activity by revealing dry surface litter from previous seasons36,48. In the period between snow melt and the onset of photosynthesis in spring, often referred to as a “spring dip”48,49, surface litter is highly flammable, leading to a peak in wildfire activity across boreal biomes48. This phenomenon has mostly been studied in North America48,49,50 and, to a lesser extent, in the European Alps24,36. However, our findings suggest that the “spring dip” and its underlying conditions are also present in snow-dominated regions across Europe, i.e., Scandinavia, the British Isles, Eastern Europe, France, and the Alps (see Figs. 2 and 3).Methodological limitations and uncertaintiesIn our study, we used the ESA FireCCI51 product to identify wildfire events and to robustly distinguish between the hydro-meteorological and land-surface conditions on wildfire start days vs days when no wildfires are burning (non-wildfire days). However, satellite-derived wildfire observations come with some limitations and uncertainties: First, wildfire observations from space, such as the ESA FireCCI51 product, have high detection uncertainties for small wildfires51,52. To exclude false detections, we only analyzed wildfires that are larger than 40 ha and have a detection confidence larger than 0.6. We acknowledge that these thresholds might have limited the sample size of wildfire events in regions with smaller wildfires, and we point out that our results related to wildfire drivers only apply to larger wildfires, i.e., those ≥40 ha. It remains to be analyzed if the importance of wildfire drivers changes with decreasing wildfire size. Second, we observe distinctive discontinuities in cumulative burned area along several country borders, e.g., Belarus and Ukraine, Russia (i.e., Kaliningrad), in the Alps on the Italian border of France and in the Pyrenees on the French border of Spain (see Fig. 1). These differences in wildfire event properties most likely originate from differences in wildfire management, forest structure and land ownership, which differ across national or even sub-national boundaries53,54,55. Third, satellite-derived wildfire products cannot account for the cause of fires, and therefore do not differentiate between prescribed burns vs wildfires. Distinguishing between human and lightning-ignited wildfires would enable accounting for the seasonality of lightning strikes and show that hydro-meteorological preconditions needed for wildfires differ by their ignition source56.Fourth, heat sources from volcanoes and gas flaring can lead to active wildfire detections in satellite-based wildfire products57. However, because the ESA FireCCI51 product incorporates a hybrid approach by considering vegetation changes in their burned area detection39, and we filtered for wildfire detections on natural land cover types only (see “FireCCI51 and derived wildfire events” under the “Methods” section), we are confident that our wildfire event dataset only includes a small number of wrongly detected wildfire events.In this study, we used random forest models and their feature importance output to understand drivers of wildfire events. As feature importance estimates can be biased when predictor variables are correlated58, we excluded highly correlated variables from the random forest model setup by using VPD and dropping relative humidity. For the variables used in the random forest model setup, we find that multiple variables have a strong correlation with temperature (see Supplementary Material Fig. S5): VPD and soil moisture have the strongest correlation with temperature, but also GPP-1M in spring and fall correlate strongly with temperature and VPD (i.e., R2 ≥ 0.5, see Supplementary Material Fig. S5). VPD and SPEI-3M are compound variables that, by definition, depend on temperature and therefore might better predict burned area than non-compound variables. The high correlation of multiple variables with temperature explains the lower ranking of temperature as a driver of wildfire events in comparison to other drivers. However, Fig. 2 shows that VPD and drought-related variables (i.e., SPEI-3M, soil moisture) are more strongly anomalous on wildfire start-days in comparison to non-wildfire days than temperature. Therefore, the correlation of these variables with temperature might slightly influence the results of the random forest analysis. However, the core findings of our analysis—that long-term drought conditions and VPD are the most important drivers of wildfires—are robust and consistent with results of other studies focusing on wildfire drivers [e.g., refs. 23,34,59,60].ConclusionWe observe wildfire-favoring conditions not only in summer and in Southern Europe, which has been the focus of many previous studies, but also in all other seasons and regions of Europe. While long-term dry conditions, described by the SPEI-3M and soil moisture deficit, and atmospheric dryness, described by VPD, are the strongest and most important drivers of wildfires, across seasons and regions, the interactions between these variables and GPP-1M vary between Southern and Northern Europe. We find that GPP-1M deficits play a major role in summer in Southern Europe, while the preconditions of summer wildfires in Northern Europe are described by low SPEI-3M and soil moisture deficits along with GPP-1M surpluses (see Figs. 2 and 3). This points to the increasing likelihood of wildfire occurrence under warmer climate conditions as prolonged drought conditions create flammable conditions in all regions and seasons of Europe (see Figs. 3 and 4 and Supplementary Material Fig. S2), as exemplified in the 2018 regional fires that occurred during drought conditions in Central and Northern Europe10.Wildfire occurrence is often the result of moderate but compounding anomalies of complex hydro-meteorological and land-surface drivers (see Fig. 2). Therefore, it is highly prone to changes in climate that affect temperature, wind speed, precipitation deficits, and VPD. Results from studies relying on future projections of the FWI show that fire danger and season length increase significantly in various regions of Europe15,17,61,62. These increases, as well as observations of wildfire occurrences outside of the typical wildfire season, such as winter wildfires in colder climates—i.e., Scandinavia, Eastern Europe, and the mountain regions of the Alps—must be considered in developing wildfire preparedness measures and planning of wildfire fighting resources. In the summer, wildfire awareness and preparedness is generally high, while the danger for wildfires in winter is more variable as wildfires in this season are driven by conditions that develop on sub-monthly timescales (see Fig. 4c). Our results suggest to consider wildfire risk in all seasons and regions of Europe, and to implement suitable adaptation measures that range from creating awareness and wildfire targeted forest management to investing in additional firefighting resources53.In summary, wildfire drivers are manifoldly affected by changing climate conditions. The likelihood of wildfires in each season and region in Europe is promoted by different combinations of emerging hydro-meteorological and land-surface drivers over multiple timescales. Our findings illustrate the regional and seasonal key drivers of wildfires and reveal their multivariate nature. Many of these drivers will be amplified under climate change, which increases wildfire risk and therefore calls for the development and implementation of effective measures to prevent and prepare for wildfires.MethodsStudy regionIn this study, we focus on the European continent, which is characterized by an arid climate with dry summers in the South, a temperate climate with warm summers in the central and western parts, a cold climate with warm and dry summers in the East, and a cold climate in the North12. We split our study area into eight different climate regions: British Isles (BI), Scandinavia (SC), France (FR), Mid-Europe (ME), Eastern Europe (EA), Alps (AL), Iberian Peninsula (IP), and Mediterranean (MD) following the PRUDENCE project37, which accounts for the north-south and east-west climate gradients of Europe. A regionalization just considering a north-south gradient, such as the SREX regions63, would not reflect the transition from an Atlantic-influenced to a continentally influenced climate within the study domain.Most of the PRUDENCE regions encompass mountain ranges, which are climatologically different from the rest of the region. Therefore, we specifically analyze the wildfire seasonality and drivers for mountain and non-mountain regions. To distinguish between mountain and non-mountain regions, we use the Global Mountains K3 dataset of ref. 64, because it was specifically developed to study global ecosystems. The Global Mountains K3 dataset was derived from the Global multi-resolution terrain elevation (GMTED2010)65 dataset at 250 m resolution64, and differentiates between high, scattered high, low, and scattered low mountains. Here, we combine high and scattered high mountains to derive a mask of high mountain regions (see “FireCCI51 and derived wildfire events” under the “Methods” section).FireCCI51 and derived wildfire eventsWe use the daily burned area from ESA FireCCI version 5.1 (FireCCI51) based on MODIS for the time period 2001–202039. The algorithm of the FireCCI51 product is based on a hybrid approach of active wildfire and burned area detection39. The thermal information of active wildfire detections is used to derive the start dates of fires, and change detection is used to classify the burn status of a pixel39.We derive wildfire events for our analysis based on the daily pixel products of ESA’s FireCCI51 product at 250 m. For each pixel and day in the 20-year time period, we filter for natural vegetation land cover classes that do not include croplands or flooded areas at any fraction (see land cover categories 50–150 in Annex 1 in ref. 66). Then, we apply a filter on the detection confidence and select only wildfire detections with a confidence level larger than 0.6 to reduce potential false-positive wildfire detections from reflectance-based confusion due to sunglint57,67, cloud shadows, or angular effects51. We then create daily binary masks of wildfire (non-)occurrences, where 1 marks wildfire pixels and 0 marks non-wildfire pixels. These daily binary masks are resampled to approximately 5.5 km resolution, which represents the grid cell size of the reanalysis dataset CERRA (see section “Hydro-meteorological and land-surface drivers” under the “Methods” section), by using the count of burned pixels on a given day. After resampling, we remap the coarsened 5.5 km grid from geographical coordinates to the projection of the CERRA dataset, i.e., the Lambert Conformal Conic projection, by applying a first-order conservative remapping68 (CDO version 1.9.669). Based on the number of originally burned pixels, we derive burned area in hectares by multiplying the number of burned pixels by the original pixel area at the given latitude for the 0.0022457331 degree grid cell resolution of the FireCCI51 pixel product66, e.g., approximately 250 × 250 m = 62500 m2 at the equator. Similar to the FireCCI51 grid product at 0.25 degree (30 km) resolution (see FireCCI51 Product User Guide66), we assume that an original pixel is 100% burned when aggregating the original grid cells to the CERRA resolution of 5.5 km.From this 5.5 km daily observation-based product, we identify wildfire events in two steps and pixel-by-pixel: First, we drop all wildfire pixels that belong to fires smaller than 40 ha because detecting small fires is challenging and uncertain39. We use a 40 ha threshold as a compromise between threshold values suggested in the studies of refs. 40,45,52,70, 52 explain that, depending on the satellite overpass angle, the minimum detectable wildfire size of active fires increases by a factor of three (i.e., 39 ha for the MODIS active fire product52). While the minimum fire size thresholds suggested in refs. 40,45,70 vary between 21 and 100 ha, the FireCCI51 product has a higher resolution than the MODIS active fire product, allowing us to use a smaller threshold. We therefore use a threshold of 40 ha for minimum fire size, which compromises between detection uncertainties and ensuring that we analyze wildfires of impact. Second, we identify the start dates of wildfire events in each pixel by applying a right-aligned rolling window of 4 days that derives the cumulative sum of burned area over the time window. We use a 4-day time window to avoid the fragmentation of large fires into multiple events and to merge smaller fires into one event. This 4-day threshold is consistent with the original FireCCI51 algorithm and has been shown to be a good compromise between shorter time steps that would lead to more fragmented fire events and longer time steps that would lead to fewer but larger fire events71. The rolling cumulative burned area indicates whether a wildfire is active for multiple days. As long as the cumulative sum of burned area in the right-aligned rolling window is larger than zero, the last day of the wildfire is still within the considered time window. Whenever the rolling sum in a pixel reaches zero again, we subtract 3 days to identify the end date of the wildfire event to account for the 3 additional days in the rolling window. Third, based on the identified start and end dates, we derive the total burned area of each wildfire event in each pixel as the sum of burned area between the start and end dates. Fourth, we assign an event-label to each wildfire event in each pixel, based on the start date and location, represented by the 5.5 km cell’s centroid. Lastly, we split this dataset into (a) days on which wildfire events start, (b) days on which wildfires burn, including the start date, and (c) days with no wildfires starting or burning, which corresponds to all days not included in (b). In our analyses and presented results, we use subset (a) to describe conditions on the days when wildfires start and subset (c) for conditions on non-wildfire days.Hydro-meteorological and land-surface driversFor the analysis of hydro-meteorological and land-surface drivers of wildfires, we use three data products: (1) the Copernicus European Regional Reanalysis dataset (CERRA)72 for daily temperature (mean and maximum; 2-m), 10-m surface wind speed, relative humidity (2-m), evaporation, snow depth, volumetric soil moisture, and radiation, (2) the CERRA-Land dataset73 for daily precipitation, and (3) the MODIS’ MOD17A2H product74 for GPP, which is available at a temporal resolution of 8 days. To make the values and value ranges of all variables comparable with each other, we standardize all variables to unit variance (mean of 0 and SD of 1 following a normal distribution) by using their empirical distribution function (see “Methods”—“Standardization of wildfire driving variables” section). We provide an overview of the variables, their data sources, and their standardization method in Table 1.Table 1 Variables used to derive the hydro-meteorological and land-surface drivers of wildfires in the analysisFull size tableCERRA and CERRA-Land are high-resolution (5.5 km), deterministic reanalysis products that cover the domain of Europe for the time period 1984–2021 at 3-hourly temporal resolution75 and are driven by lateral boundary conditions from ERA-572,76. We derive precipitation from CERRA-Land because it uses additional quality-controlled precipitation observations in comparison to CERRA. CERRA and CERRA-Land show a better representation of most climate variables at surface levels and also capture local extreme events better than ERA-575,77. Based on the original variables from CERRA, we derive VPD78, potential evapotranspiration (PET)78, and the three-month Standardized Evapotranspiration Precipitation Index (SPEI-3M)79, which are described in detail in the supplementary material chapter Supplementary Methods.We add GPP data from MODIS (i.e., MOD17A2H v006 product74) to our variable collection to account for fuel availability and interactions between hydro-meteorological and land-surface conditions. GPP describes the sum of daily photosynthesis74 and is used as an indicator for available biomass to burn in this study. The MOD17A2H product estimates GPP by combining the measured fraction of absorbed photosynthetically active radiation (FPAR) with additional data sources for photosynthetically active radiation (PAR), i.e., meteorological fields from GMAO/NASA80, and biome-specific parameters for radiation use efficiency, which describes how efficiently absorbed radiation is converted to vegetation productivity74. The product is available as 8-day composites at 500 m spatial resolution, which represent cumulative GPP for the 8-day period74. To match the spatial resolution of the GPP product with that of the other driver variables, we resample and regrid the 500 m MODIS grid to the CERRA reanalysis resolution and projection in the same manner as we processed the FireCCI51 250 m dataset (for details, see “FireCCI51 and derived wildfire events” under the “Methods” section). To account for fluctuations in fuel availability, i.e., available biomass to burn, we apply a rolling window of 4 to the 8-day cumulative GPP from the MOD17A2H product to capture monthly GPP summations (GPP-1M). This rolling window summarizes four times 8-day GPP summations and therefore, represents cumulative GPP for 32 days, which we consider to reflect one month at each 8-day observation point. While GPP-1M is not a direct measure of fuel load, monthly sums allow us to capture moisture losses in biomass, which are relevant for fire activity26,81,82. To match the 8-day temporal resolution of the GPP-1M dataset with the daily resolution of the other variables, we apply a linear interpolation to fill in the days between the 8-day GPP-1M values, after the standardization (see “FireCCI51 and derived wildfire events” under the “Methods” section).Spatial and temporal patterns of wildfiresWe evaluate spatial and temporal patterns of wildfires for each region, its respective mountain and non-mountain regions, and for four different seasons, i.e., spring (MAM), summer (JJA), fall (SON), and winter (DJF). To study spatial patterns, we derive the cumulative burned area between 2001 and 2020 of each CERRA grid cell. To investigate seasonal patterns, we derive the seasonal fraction of total annual burned area and number of wildfires separately for the mountain and non-mountain regions in each region.Standardization of wildfire driving variablesWe standardize all driver variables, i.e., maximum temperature, vapor pressure deficit (VPD), 10-m surface wind speed, GPP, precipitation, soil moisture, and snow depth, to unit variance using their empirical distribution83. We do not standardize the SPEI-3M because it is already a standardized index79. Through the standardization, each daily value of the 20-year time series (2001–2020) is expressed as an SD from the climate mean, which equals zero. We apply the standardization pixel-by-pixel for each CERRA grid cell to account for local climate conditions and to be able to compare the results across regions.The standardization based on the empirical distribution is applied in two steps: First, we estimate the empirical cumulative distribution function F(x) by estimating the percentile rank of each value P(x). The probabilities in F(x) and the respective percentile scores P(x) are based on the rank given by the indicator function ({mathbb{1}}) (Xi ≤ x). The indicator function describes the estimation of the rank by being equal to 1 if Xi ≤ x and equal to 0 if Xi ≥ x for the given number of observations (n):$$P(x)=F(x)=frac{1}{n}{sum }_{i = 1}^{n}{mathbb{1}}({X}_{i}le x).$$
    (1)
    For variables, which have minimum values of zero on the left side of their distribution (i.e., VPD, precipitation, snow depth; see Table 1), we randomize the variables’ zero values with a very small μ (i.e., 0.0e-27) following a normal distribution. The goal of this step is to avoid the over-representation of zeros in the value ranking. Second, we map the percentile scores P(x) to the quantiles of a normal distribution z(x), by using the inverse of the cumulative distribution function (Φ−1):$$z(x)={{{Phi }}}^{-1}(P(x)).$$
    (2)
    We evaluate the standardized wildfire driving variables for the start date of wildfire events and compare them with the conditions of non-wildfire days. To derive two distinct datasets, we sample the drivers for the start dates of wildfires and days when no wildfires occur, i.e., datasets (a) and (c) described in “FireCCI51 and derived wildfire events” under the “Methods” section. Based on these two samples, we test for independence (p-value ≤ 0.05) using the non-parametric Mann–Whitney U-test84 to show significant differences in hydro-meteorological and land-surface preconditions between wildfire start days in comparison to non-wildfire days.Variable importance of wildfire driversWe use random forest models85 to quantify the importance of different hydro-meteorological and land-surface drivers for the occurrence of wildfire events. Random forest models represent a supervised machine learning algorithm, which creates multiple decision trees by randomly sampling training observations for a classification problem (in this case, wildfire vs non-wildfire events). The resulting model represents the average of all decision trees and calculates the variable importance based on the mean decrease in impurity for the different tree splits, also known as Gini importance85.Here, we set up multiple random forest models for each region and season using maximum temperature, VPD, 10-m surface wind speed, GPP (cumulative over 32 days), the SPEI-3M, soil moisture, and snow depth as predictor variables and the binary flag of a wildfire event start as a target variable. We use the Python package SCIKIT-LEARN V.1.3.086 and its default hyper-parameters (see https://scikit-learn.org/stable/modules/generated/sklearn.ensemble.RandomForestClassifier.html), as hyper-parameter tuning does not necessarily improve prediction skill87. First, we randomly select non-wildfire days in the respective season and region that match the number of observed wildfire days to account for dataset imbalance. Second, we perform a five-fold cross-validation by training the model on 80% of these samples and evaluating it on the remaining 20%. To capture the variability of the random-forest decision trees, we repeat the previous steps 100 times. The presented results on feature importance show the mean importance for each region and season across the 100 random forest models.We evaluate model performance using the overall accuracy of the confusion matrix, which shows the ratio between correct classifications (true positive and true negative) and all classifications88. The mean over all accuracies for 100 random forest models in each region and season ranges between 0.76 and 0.96 (see Supplementary Material Fig. S2). We observe the highest accuracies in winter in all regions, except the Mediterranean and the Iberian Peninsula. In these two regions, we also observe the overall lowest accuracies, which occur in summer and fall, when conditions are strongly wildfire-favoring, also on days with no wildfire occurrences. Given the high overall accuracies (see Supplementary Material Fig. S5), we expect to derive meaningful findings in terms of the variable importance of wildfire drivers.Time of precondition emergence of wildfire driversWe study how long wildfire preconditions have been anomalous before wildfire events, by calculating the rolling mean of the standardized driving variables for different time steps. We increase the window sizes of the right-aligned rolling mean iteratively from daily levels (i.e., 0 (day of wildfire-start), 3, 7, 14 days before wildfires ignite) to monthly levels (1–12 months, where we count 30 days for each month) up to one year (i.e., 360 days) before the start day of each wildfire event for all variables. We assume that the incrementally increasing window size indicates how long anomalous conditions of wildfire driving preconditions persist prior to a wildfire event. To set this information in context with non-wildfire conditions, we derive the same rolling mean for all days on which no wildfires occur (see non-wildfire dataset (c) in “FireCCI51 and derived wildfire events” under “Methods” section). We conduct the precondition analysis on the wildfire event scale and then average the results for each region and season. This results in two graphs, one for preconditions on wildfire start days and one for those on non-wildfire days (see Fig. 4a), for each region, season, and variable. In the next step, we identify the time point where these two graphs deviate from each other. Specifically, we derive the difference between the graph for wildfire start days and the graph for non-wildfire days and use different SD levels, i.e., 1 SD, 1.5 SD, and 2 SD, of this difference as a buffer for natural variability on the non-wildfire graph. Last, we identify the time step, where the anomalies of the wildfire days emerge from these SD buffers around the non-wildfire anomalies, and call this time step “time of precondition emergence” (ToPE). We find that the 1 SD threshold coincides consistently with the onset of the divergence between wildfire and non-wildfire preconditions across variables and regions (see Supplementary Material Fig. S4). Therefore, we use ToPE based on one SD in our analysis. We use +1 SD for maximum temperature, VPD, and 10-m surface wind speed, and −1 SD for SPEI-3M, soil moisture, and snow depth. For GPP (GPP-1M), we find that in spring and summer GPP-1M is more abundant on non-wildfire days, whereas in fall and winter GPP-1M is more abundant on wildfire-days compared to non-wildfire days (see Fig. 2), which is why we derive the lower level of ToPE (−1 SD) for spring and summer and the upper anomaly persistence (+1 SD) for fall and winter.Reporting summaryFurther information on research design is available in the Nature Portfolio Reporting Summary linked to this article.

    Data availability

    FireCCI data can be downloaded from the Copernicus repository https://doi.org/10.24381/cds.f333cf85. CERRA data can be downloaded from the Copernicus repository https://doi.org/10.24381/cds.622a565a. CERRA-Land data can be downloaded from the Copernicus repository https://doi.org/10.24381/cds.a7f3cd0b. MODIS Gross Primary Productivity can be downloaded from https://www.earthdata.nasa.gov/data/catalog/lpcloud-mod17a2h-061/. The mountain region mask is available at https://data.usgs.gov/datacatalog/data/USGS:638fbf72d34ed907bf7d3080. We provide the derived wildfire events and drivers, which we used in the analysis, through a dataset repository on Envidat (https://doi.org/10.16904/envidat.575).
    Code availability

    The code to reproduce the analysis and the figures is available on GitLab: (https://gitlab.com/jumi26/wildfire-drivers-in-europe.git).
    ReferencesSan-Miguel-Ayanz, J. et al. Advance report on Forest Fires in Europe, Middle East and North Africa 2023 (European Commission, 2024).San-Miguel-Ayanz, J. et al. Advance Report on Forest Fires in Europe, Middle East and North Africa 2022 (European Commission, 2023).Turco, M. et al. Decreasing fires in Mediterranean Europe. PLOS One 11, e0150663 (2016).Article 

    Google Scholar 
    Jones, M. W. et al. Global and regional trends and drivers of fire under climate change. Rev. Geophys. 60, e2020RG000726 (2022).Article 

    Google Scholar 
    Andela, N. et al. A human-driven decline in global burned area. Science 356, 1356–1362 (2017).Article 
    CAS 

    Google Scholar 
    Grünig, M., Seidl, R. & Senf, C. Increasing aridity causes larger and more severe forest fires across Europe. Glob. Change Biol. 29, 1648–1659 (2023).Article 

    Google Scholar 
    Grau-Andrés, R., Moreira, B. & Pausas, J. G. Global plant responses to intensified fire regimes. Glob. Ecol. Biogeogr. 33, e13858 (2024).Article 

    Google Scholar 
    Pausas, J. G. & Paula, S. Fuel shapes the fire-climate relationship: evidence from Mediterranean ecosystems. Glob. Ecol. Biogeogr. 21, 1074–1082 (2012).Article 

    Google Scholar 
    Bastos, A. et al. Impacts of extreme summers on european ecosystems: a comparative analysis of 2003, 2010 and 2018. Philos. Trans. R. Soc. B Biol. Sci. 375, 20190507 (2020).Article 
    CAS 

    Google Scholar 
    San-Miguel-Ayanz, J. et al. Forest Fires in Europe, Middle East and North Africa 2018 (European Commission, 2019).Bevacqua, E. et al. Direct and lagged climate change effects intensified the 2022 European drought. Nat. Geosci. 17, 1–8 (2024).Beck, H. E. et al. Present and future köppen-geiger climate classification maps at 1-km resolution. Sci. Data 5, 180214 (2018).Article 

    Google Scholar 
    Felsche, E., Böhnisch, A., Poschlod, B. & Ludwig, R. European hot and dry summers are projected to become more frequent and expand northwards. Commun. Earth Environ. 5, 410 (2024).Article 

    Google Scholar 
    Arnell, N. W., Freeman, A. & Gazzard, R. The effect of climate change on indicators of fire danger in the UK. Environ. Res. Lett. 16, 44027 (2021).Article 

    Google Scholar 
    Miller, J., Böhnisch, A., Ludwig, R. & Brunner, M. I. Climate change impacts on regional fire weather in heterogeneous landscapes of Central Europe. Nat. Hazards Earth Syst. Sci. 24, 1–25 (2023).De Rigo, D., Libertá, G., Houston Durrant, T., Artés Vivancos, T. & San-Miguel-Ayanz, J. Forest fire Danger Extremes in Europe Under Climate Change: Variability and Uncertainty (Publications Office of the European Union, 2017).El Garroussi, S., Di Giuseppe, F., Barnard, C. & Wetterhall, F. Europe faces up to tenfold increase in extreme fires in a warming climate. npj Clim. Atmos. Sci. 7, 30 (2024).Article 

    Google Scholar 
    Zscheischler, J. et al. A typology of compound weather and climate events. Nat. Rev. Earth Environ. 1, 333–347 (2020).Article 

    Google Scholar 
    Van Wagner, C. E. Development and structure of the Canadian Forest Fire Weather Index System (Canadian Forestry Service, 1987); https://cfs.nrcan.gc.ca/publications?id=19927.Grillakis, M. et al. Climate drivers of global wildfire burned area. Environ. Res. Lett. 17, 045021 (2022).Article 

    Google Scholar 
    Jain, P. et al. Drivers and impacts of the record-breaking 2023 Wildfire Season in Canada. Nat. Commun. 15, 6764 (2024).Article 
    CAS 

    Google Scholar 
    Giuseppe, F. D. Accounting for fuel in fire danger forecasts: the fire occurrence probability index (FOPI). Environ. Res. Lett. 18, 064029 (2023).Article 

    Google Scholar 
    Turco, M. et al. On the key role of droughts in the dynamics of summer fires in Mediterranean Europe. Sci. Rep. 7, 81 (2017).Article 

    Google Scholar 
    Conedera, M., Feusi, J., Pezzatti, G. B. & Krebs, P. Linking the future likelihood of large fires to occur on mountain slopes with fuel connectivity and topography. Nat. Hazards 120, 1–17 (2024).McCarty, J. L., Smith, T. E. L. & Turetsky, M. R. Arctic fires re-emerging. Nat. Geosci. 13, 658–660 (2020).Article 
    CAS 

    Google Scholar 
    Ermitão, T., Gouveia, C. M., Bastos, A. & Russo, A. C. Interactions between hot and dry fuel conditions and vegetation dynamics in the 2017 fire season in Portugal. Environ. Res. Lett. 17, 095009 (2022).Article 

    Google Scholar 
    Rodrigues, M., Trigo, R. M., Vega-García, C. & Cardil, A. Identifying large fire weather typologies in the Iberian Peninsula. Agric. For. Meteorol. 280, 107789 (2020).Article 

    Google Scholar 
    Torres-Vázquez, M. N. et al. Large increase in extreme fire weather synchronicity over Europe. Environ. Res. Lett. 20, 024045 (2025).Article 

    Google Scholar 
    Pausas, J. G. Pyrogeography across the western Palaearctic: a diversity of fire regimes. Glob. Ecol. Biogeogr. 31, 1923–1932 (2022).Article 

    Google Scholar 
    Gincheva, A. et al. The interannual variability of global burned area is mostly explained by climatic drivers. Earth’s. Future 12, e2023EF004334 (2024).Article 

    Google Scholar 
    McNorton, J., Moreno, A., Turco, M., Keune, J. & Di Giuseppe, F. Hydroclimatic rebound drives extreme fire in california’s non-forested ecosystems. Glob. Change Biol. 31, e70481 (2025).Article 
    CAS 

    Google Scholar 
    Swain, D. L. et al. Increasing hydroclimatic whiplash can amplify wildfire risk in a warming climate. Glob. Change Biol. 31, e70075 (2025).Article 
    CAS 

    Google Scholar 
    Swain, D. L. et al. Hydroclimate volatility on a warming Earth. Nat. Rev. Earth Environ. 6, 35–50 (2025).Article 

    Google Scholar 
    Kuhn-Régnier, A. et al. The importance of antecedent vegetation and drought conditions as global drivers of burnt area. Biogeosciences 18, 3861–3879 (2021).Article 

    Google Scholar 
    Krikken, F., Lehner, F., Haustein, K., Drobyshev, I. & van Oldenborgh, G. J. Attribution of the role of climate change in the forest fires in Sweden 2018. Nat. Hazards Earth Syst. Sci. 21, 2169–2179 (2021).Article 

    Google Scholar 
    Conedera, M. et al. Characterizing Alpine pyrogeography from fire statistics. Appl. Geogr. 98, 87–99 (2018).Article 

    Google Scholar 
    Christensen, J. H. & Christensen, O. B. A summary of the PRUDENCE model projections of changes in European climate by the end of this century. Clim. Change 81, 7–30 (2007).Article 

    Google Scholar 
    Karagulle, D. et al. Modeling global Hammond landform regions from 250-m elevation data. Trans. GIS 21, 1040–1060 (2017).Article 

    Google Scholar 
    Lizundia-Loiola, J., Otón, G., Ramo, R. & Chuvieco, E. A spatio-temporal active-fire clustering approach for global burned area mapping at 250 m from MODIS data. Remote Sens. Environ. 236, 111493 (2020).Article 

    Google Scholar 
    Khorshidi, M. S. et al. Increasing concurrence of wildfire drivers tripled megafire critical danger days in Southern California between 1982 and 2018. Environ. Res. Lett. 15, 104002 (2020).Article 

    Google Scholar 
    Di Giuseppe, F., McNorton, J., Lombardi, A. & Wetterhall, F. Global data-driven prediction of fire activity. Nat. Commun. 16, 2918 (2025).Article 

    Google Scholar 
    McNorton, J. R. & Di Giuseppe, F. A global fuel characteristic model and dataset for wildfire prediction. Biogeosciences 21, 279–300 (2024).Article 

    Google Scholar 
    Fu, Z. et al. Atmospheric dryness reduces photosynthesis along a large range of soil water deficits. Nat. Commun. 13, 989 (2022).Article 
    CAS 

    Google Scholar 
    Bastos, A. et al. Direct and seasonal legacy effects of the 2018 heat wave and drought on European ecosystem productivity. Sci. Adv. 6, eaba2724 (2020).Article 
    CAS 

    Google Scholar 
    Ruffault, J. et al. Increased likelihood of heat-induced large wildfires in the Mediterranean. Sci. Rep. 10, 13790 (2020).Article 
    CAS 

    Google Scholar 
    Ruffault, J., Moron, V., Trigo, R. M. & Curt, T. Daily synoptic conditions associated with large fire occurrence in Mediterranean France: evidence for a wind-driven fire regime. Int. J. Climatol. 37, 524–533 (2017).Article 

    Google Scholar 
    Giannaros, T. M. et al. Meteorological analysis of the 2021 extreme wildfires in Greece: lessons learned and implications for early warning of the potential for pyroconvection. Atmosphere 13, 475 (2022).Parisien, M.-A., Barber, Q. E., Flannigan, M. D. & Jain, P. Broadleaf tree phenology and springtime wildfire occurrence in boreal Canada. Glob. Change Biol. 29, 6106–6119 (2023).Article 
    CAS 

    Google Scholar 
    Jolly, W. et al. Seasonal variations in red pine (Pinus resinosa) and jack pine (Pinus banksiana) foliar physio-chemistry and their potential influence on stand-scale wildland fire behavior. For. Ecol. Manag. 373, 167–178 (2016).Article 

    Google Scholar 
    Westerling, A. L. Increasing western US forest wildfire activity: sensitivity to changes in the timing of spring. Philos. Trans. R. Soc. B Biol. Sci. 371, 20150178 (2016).Article 

    Google Scholar 
    Chuvieco, E. et al. Generation and analysis of a new global burned area product based on MODIS 250m reflectance bands and thermal anomalies. Earth Syst. Sci. Data 10, 2015–2031 (2018).Article 

    Google Scholar 
    Giglio, L., Loboda, T., Roy, D. P., Quayle, B. & Justice, C. O. An active-fire based burned area mapping algorithm for the MODIS sensor. Remote Sens. Environ. 113, 408–420 (2009).Article 

    Google Scholar 
    Pandey, P. et al. A global outlook on increasing wildfire risk: Current policy situation and future pathways. Trees For. People 14, 100431 (2023).Article 

    Google Scholar 
    Tedim, F. et al. Forest fire causes and motivations in Southern and South-Eastern Europe through the perception of experts: contribution to enhance the current policies. Forests 13, 562 (2022).San-Miguel-Ayanz, J., Durrant, T. & Camia, A. The European Fire Database: Technical Specifications and Data Submission: Executive Report (Institute for Environment and Sustainability) (Joint Research Centre, LU, 2014).Balch, J. K. et al. Human-started wildfires expand the fire niche across the United States. Proc. Natl. Acad. Sci. USA 114, 2946–2951 (2017).Article 
    CAS 

    Google Scholar 
    Wooster, M. J. et al. Satellite remote sensing of active fires: history and current status, applications and future requirements. Remote Sens. Environ. 267, 112694 (2021).Article 

    Google Scholar 
    Strobl, C., Boulesteix, A.-L., Kneib, T., Augustin, T. & Zeileis, A. Conditional variable importance for random forests. BMC Bioinforma. 9, 307 (2008).Article 

    Google Scholar 
    Qiu, L., Chen, J., Fan, L., Sun, L. & Zheng, C. High-resolution mapping of wildfire drivers in California based on machine learning. Sci. Total Environ. 833, 155155 (2022).Article 
    CAS 

    Google Scholar 
    Rodrigues, M., Resco de Dios, V., Sil, n, Cunill Camprubí, n & Fernandes, P. M. VPD-based models of dead fine fuel moisture provide best estimates in a global dataset. Agric. For. Meteorol. 346, 109868 (2024).Article 

    Google Scholar 
    Hetzer, J., Forrest, M., Ribalaygua, J., Prado-López, C. & Hickler, T. The fire weather in Europe: large-scale trends towards higher danger. Environ. Res. Lett. 19, 084017 (2024).Article 

    Google Scholar 
    Quilcaille, Y., Batibeniz, F., Ribeiro, A. F. S., Padrón, R. S. & Seneviratne, S. I. Fire weather index data under historical and shared socioeconomic pathway projections in the 6th phase of the Coupled Model Intercomparison Project from 1850 to 2100. Earth Syst. Sci. Data 15, 2153–2177 (2023).Article 

    Google Scholar 
    Seneviratne, S. I. et al. Changes in climate extremes and their impacts on the natural physical environment. In Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation (eds Field, C. B., Barros, V., Stocker, T. F. & Dahe, Q.) 109–230 (Cambridge University Press, 2012).Sayre, R. et al. A new high-resolution map of world mountains and an online tool for visualizing and comparing characterizations of global mountain distributions. Mt. Res. Dev. 38, 240–249 (2018).Article 

    Google Scholar 
    Danielson, J. J. & Gesch, D. B. Global Multi-resolution Terrain Elevation Data 2010 (GMTED2010). Technical Report 2011-1073 (U.S. Geological Survey, 2011).Pettinari, M. L., Lizundia-Loiola, J. & Chuvieco, E. ESA Climate Change Initiative—FireCCI D4.2.1 Product User Guide—MODIS (PUG). Technical Report (Fire CCI, 2021).Justice, C. O. et al. The MODIS fire products. Remote Sens. Environ. 83, 244–262 (2002).Article 

    Google Scholar 
    Jones, P. W. First- and second-order conservative remapping schemes for grids in spherical coordinates. Monthly Weather Rev. 127, 2204–2210 (1999).Schulzweida, U. CDO User Guide. Version number: 2.3.0. https://zenodo.org/doi/10.5281/zenodo.10020800 (2023).Randerson, J. T., Chen, Y., van der Werf, G. R., Rogers, B. M. & Morton, D. C. Global burned area and biomass burning emissions from small fires. J. Geophys. Res. Biogeosci. 117, G04012 (2012).Lizundia-Loiola, J., Franquesa, M., Khairoun, A. & Chuvieco, E. Global burned area mapping from Sentinel-3 synergy and VIIRS active fires. Remote Sens. Environ. 282, 113298 (2022).Article 

    Google Scholar 
    Schimanke, S. et al. CERRA Sub-daily Regional Reanalysis Data for Europe on Single Levels from 1984 to Present (European Union, 2021).Verrelle, A. et al. CERRA-Land Sub-daily Regional Reanalysis Data for Europe from 1984 to Present (European Union, 2022).Running, S., Mu, Q. & Zhao, M. MOD17A2H MODIS/Terra Gross Primary Productivity 8-Day L4 Global 500m SIN Grid V006 (NASA, 2015).Ridal, M. et al. CERRA, the Copernicus European Regional Reanalysis system. Q. J. R. Meteorol. Soc. 150, 3385–3411 (2024).Article 

    Google Scholar 
    Hersbach, H. et al. The ERA5 global reanalysis. Q. J. R. Meteorol. Soc. 146, 1999–2049 (2020).Article 

    Google Scholar 
    Wood, R. R. et al. Comparison of high-resolution climate reanalysis datasets for hydro-climatic impact studies. Hydrol. Earth Syst. Sci. 29, 4153–4178 (2025).Article 

    Google Scholar 
    Allen, R. G., Pereira, L. S. & Raes, D. (eds.) Crop Evapotranspiration: Guidelines for Computing Crop Water Requirements 56 (Food and Agriculture Organization of the United Nations, Rome, 1998).Vicente-Serrano, S. M., Beguería, S. & López-Moreno, J. I. A multiscalar drought index sensitive to global warming: the standardized precipitation evapotranspiration index. J. Clim. 23, 1696–1718 (2010).Article 

    Google Scholar 
    Rienecker, M. M. et al. The GEOS-5 Data Assimilation System-Documentation of Versions 5.0.1, 5.1.0, and 5.2.0. Technical Report NASA/TM-2008-104606-VOL-27 (NASA, 2008).Forrest, M. et al. Understanding and simulating cropland and non-cropland burning in Europe using the BASE (Burnt Area Simulator for Europe) model. Biogeosciences 21, 5539–5560 (2024).Article 

    Google Scholar 
    Lian, X., Li, Y., Liu, J., Kornhuber, K. & Gentine, P. Northern ecosystem productivity reduced by Rossby-wave-driven hot-dry conditions. Nat. Geosci. 18, 1–9 (2025).Allen, S. & Otero, N. Standardised indices to monitor energy droughts. Renew. Energy 217, 119206 (2023).Article 

    Google Scholar 
    Mann, H. B. & Whitney, D. R. On a test of whether one of two random variables is stochastically larger than the other. Ann. Math. Stat. 18, 50–60 (1947).Article 

    Google Scholar 
    Breiman, L. Random forests. Mach. Learn. 45, 5–32 (2001).Article 

    Google Scholar 
    Pedregosa, F. et al. Scikit-learn: machine learning in Python. J. Mach. Learn. Res. 12, 2825–2830 (2011).van Hamel, A. & Brunner, M. I. Trends and drivers of water temperature extremes in mountain rivers. Water Resour. Res. 60, e2024WR037518 (2024).Article 

    Google Scholar 
    Ma, J. et al. Analyzing the leading causes of traffic fatalities using XGBoost and grid-based analysis: a city management perspective. IEEE Access 7, 148059–148072 (2019).Article 

    Google Scholar 
    Download referencesAcknowledgementsThis research was funded through the FoFix Project by ETH Zurich. We thank the Doc-Mobility Fellowship of ETH Zurich and the Jackson School of Geosciences at the University of Texas at Austin, for supporting this collaboration.FundingOpen access funding provided by Swiss Federal Institute of Technology Zurich.Author informationAuthors and AffiliationsInstitute for Atmospheric and Climate Science, ETH Zurich, Zurich, SwitzerlandJulia Miller & Manuela I. BrunnerWSL Institute for Snow and Avalanche Research SLF, Davos Dorf, SwitzerlandJulia Miller & Manuela I. BrunnerClimate Change Extremes and Natural Hazards in Alpine Regions Research Center CERC, Davos Dorf, SwitzerlandJulia Miller & Manuela I. BrunnerJackson School of Geosciences, Institute for Geophysics, University of Austin, Austin, TX, USADanielle ToumaAuthorsJulia MillerView author publicationsSearch author on:PubMed Google ScholarDanielle ToumaView author publicationsSearch author on:PubMed Google ScholarManuela I. BrunnerView author publicationsSearch author on:PubMed Google ScholarContributionsJ.M., D.T., and M.I.B. conceptualized the study. J.M. conducted the data collection, analysis, and visualization. J.M. wrote the original draft, which was revised and edited by D.T., M.I.B., and J.M.Corresponding authorCorrespondence to
    Julia Miller.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Peer review

    Peer review information
    Communications Earth and Environment thanks the anonymous reviewers for their contribution to the peer review of this work. Peer review was double-anonymous. Primary handling editors: Erika Buscardo and Mengjie Wang. 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 Material FileReporting SummaryRights 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 articleMiller, J., Touma, D. & Brunner, M.I. Compounding preconditions of wildfires vary in time and space within Europe.
    Commun Earth Environ 6, 1005 (2025). https://doi.org/10.1038/s43247-025-02955-1Download citationReceived: 18 March 2025Accepted: 28 October 2025Published: 15 December 2025Version of record: 15 December 2025DOI: https://doi.org/10.1038/s43247-025-02955-1Share 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