More stories

  • in

    Mycorrhizal colonization of dryland tree establishment depends on soil microbial cooperation

    AbstractMycorrhizal fungi serve as fundamental agents in forest establishment and progression, underpinning critical ecosystem functions through symbiotic root associations. Drylands, which cover nearly half of Earth’s land, have limited forest establishment, and factors influencing mycorrhization in these stressful environments remain unclear. Here, we integrate large-scale field surveys along aridity gradients with greenhouse experiments and over 33,000 microscopic mycorrhizal observations, revealing that aridity significantly enhances mycorrhization. Mycorrhizal fungi undergo niche modification, whereby facilitative microbial interactions promote mycorrhization under aridity stress. We identify a core synthetic microbial community linked to mycorrhization and provide mechanistic evidence that this community facilitates mycorrhization through physical attachment to fungal hyphae and by alleviating soil metabolite inhibition that otherwise suppresses mycorrhization under arid conditions. In this work, our findings highlight the role of microbial interkingdom interactions in driving tree mycorrhizal colonization in arid regions, offering critical insights for guiding tree planting and restoration efforts in drylands.

    Similar content being viewed by others

    Aridity-driven shift in biodiversity–soil multifunctionality relationships

    Article
    Open access
    09 September 2021

    Global hotspots of mycorrhizal fungal richness are poorly protected

    Article
    Open access
    23 July 2025

    Pervasive associations between dark septate endophytic fungi with tree root and soil microbiomes across Europe

    Article
    Open access
    02 January 2024

    Data availability

    All raw plant RNA-seq data, amplicon sequencing data generated in this study have been deposited in the Sequence Read Archive (http://www.ncbi.nlm.nih.gov/sra). Raw amplicon sequences derived from the field survey are publicly available under NCBI BioProject number PRJNA1311795 (16S rRNA gene) and PRJNA1311818 (ITS). All 16S rRNA sequence data of bacterial strains are publicly available under NCBI BioProject number PRJNA1312253. The RNA-seq data for Tuber are publicly available under NCBI BioProject number PRJNA1312988. Source data are available in the Figshare database (https://doi.org/10.6084/m9.figshare.30775883). Source data are provided with this paper.
    Code availability

    Codes used in this study are available in the Figshare database (https://doi.org/10.6084/m9.figshare.30685235).
    ReferencesRen, S. et al. Storage potential of soil functional carbon fractions in the World’s largest plantations. Adv. Sci. 12, e04995 (2025).
    Google Scholar 
    Wang, Y. J. et al. Land availability and policy commitments limit global climate mitigation from forestation. Science 389, 931–934 (2025).
    Google Scholar 
    Garrett, L., Lévite, H., Besacier, C., Alekseeva, N. & Duchelle, M. The key role of forest and landscape restoration in climate action. (FAO, 2022).Reynolds, J. F. et al. Global desertification: building a science for dryland development. Science 316, 847–851 (2007).
    Google Scholar 
    Ma, L. L. et al. Planted forests in China have higher drought risk than natural forests. Glob. Change Biol. 31, e70055 (2025).
    Google Scholar 
    Koppa, A. et al. Dryland self-expansion enabled by land-atmosphere feedbacks. Science 385, 967–972 (2024).
    Google Scholar 
    Gu, L., Schumacher, D. L., Wang, H. M., Yin, J. & Fischer, E. M. Land-atmosphere feedbacks drive dryland drought and expansion under climate warming. Innovation 6, 100863 (2025).
    Google Scholar 
    Cosme, M. Mycorrhizas drive the evolution of plant adaptation to drought. Commun. Biol. 6, 346 (2023).
    Google Scholar 
    Sachsenmaier, L. et al. Forest growth resistance and resilience to the 2018–2020 drought depend on tree diversity and mycorrhizal type. J. Ecol. 112, 1787–1803 (2024).
    Google Scholar 
    Poudel, M. et al. The role of plant-associated bacteria, fungi, and viruses in drought stress mitigation. Front. Microbiol. 12, 743512 (2021).
    Google Scholar 
    Hu, M. Y. et al. Plant functional traits affect biomass responses to global change: a meta-analysis. J. Ecol. 113, 2046–2065 (2025).
    Google Scholar 
    Van Nuland, M. E. et al. Global hotspots of mycorrhizal fungal richness are poorly protected. Nature 645, 414–422 (2025).
    Google Scholar 
    Jiang, F. et al. Mycorrhizal symbioses and tree diversity in global forest communities. Sci. Adv. 11, eadt5743 (2025).
    Google Scholar 
    Berdugo, M. et al. Global ecosystem thresholds driven by aridity. Science 367, 787–790 (2020).
    Google Scholar 
    Stuart, E. K. et al. Acquisition of host-derived carbon in biomass of the ectomycorrhizal fungus Pisolithus microcarpus is correlated to fungal carbon demand and plant defences. FEMS Microbiol. Ecol. 99, fiad037 (2023).
    Google Scholar 
    Canarini, A. et al. Soil fungi remain active and invest in storage compounds during drought, independent of future climate conditions. Nat. Commun. 15, 10410 (2024).
    Google Scholar 
    Xie, L. L. et al. Variations in ectomycorrhizal exploration types parallel seedling fine root traits of two temperate tree species under extreme drought and contrasting solar radiation treatments. Plant Cell Environ. 47, 5053–5066 (2024).
    Google Scholar 
    Guarnizo, A. L., Navarro-Ródenas, A., Calvo-Polanco, M., Marqués-Gálvez, J. E. & Morte, A. A mycorrhizal helper bacterium alleviates drought stress in mycorrhizal Helianthemum almeriense plants by regulating water relations and plant hormones. Environ. Exp. Bot. 207, 105228 (2023).
    Google Scholar 
    Berrios, L. et al. Positive interactions between mycorrhizal fungi and bacteria are widespread and benefit plant growth. Curr. Biol. 33, 2878–2887.e2874 (2023).
    Google Scholar 
    Navarro-Rodenas, A., Berna, L. M., Lozano-Carrillo, C., Andrino, A. & Morte, A. Beneficial native bacteria improve survival and mycorrhization of desert truffle mycorrhizal plants in nursery conditions. Mycorrhiza 26, 769–779 (2016).
    Google Scholar 
    Jorgensen, K., Clemmensen, K. E., Wallander, H. & Lindahl, B. D. Do ectomycorrhizal exploration types reflect mycelial foraging strategies? New Phytol. 237, 576–584 (2023).
    Google Scholar 
    Durán, P. et al. Microbial interkingdom interactions in roots promote Arabidopsis survival. Cell 175, 973–983.e914 (2018).
    Google Scholar 
    Bertness, M. D. & Callaway, R. Positive interactions in communities. Trends Ecol. Evol. 9, 191–193 (1994).
    Google Scholar 
    Hammarlund, S. P. & Harcombe, W. R. Refining the stress gradient hypothesis in a microbial community. Proc. Natl. Acad. Sci. USA 116, 15760–15762 (2019).
    Google Scholar 
    Hernandez, D. J., David, A. S., Menges, E. S., Searcy, C. A. & Afkhami, M. E. Environmental stress destabilizes microbial networks. ISME J. 15, 1722–1734 (2021).
    Google Scholar 
    Gao, C. et al. Co-occurrence networks reveal more complexity than community composition in resistance and resilience of microbial communities. Nat. Commun. 13, 3867 (2022).
    Google Scholar 
    Berrios, L., Ansell, T. B., Dahlberg, P. D. & Peay, K. G. Standardizing experimental approaches to investigate interactions between bacteria and ectomycorrhizal fungi. FEMS Microbiol. Rev. 49, fuae035 (2025).
    Google Scholar 
    Hao, Z. G. et al. Thresholds in aridity and soil carbon-to-nitrogen ratio govern the accumulation of soil microbial residues. Commun. Earth Environ. 2, 236 (2021).
    Google Scholar 
    Miyamoto, Y., Maximov, T. C., Kononov, A. & Sugimoto, A. Soil propagule banks of ectomycorrhizal fungi associated with Larix cajanderi above the treeline in the Siberian Arctic. Mycoscience 63, 142–148 (2022).
    Google Scholar 
    Shemesh, H., Bruns, T. D., Peay, K. G., Kennedy, P. G. & Nguyen, N. H. Changing balance between dormancy and mortality determines the trajectory of ectomycorrhizal fungal spore longevity over a 15-yr burial experiment. New Phytol. 238, 11–15 (2023).
    Google Scholar 
    Williams, A., Sinanaj, B. & Hoysted, G. A. Plant-microbe interactions through a lens: tales from the mycorrhizosphere. Ann. Bot. 133, 399–412 (2024).
    Google Scholar 
    Hoek, T. A. et al. Resource availability modulates the cooperative and competitive nature of a microbial cross-feeding mutualism. PLoS Biol. 14, e1002540 (2016).
    Google Scholar 
    Piccardi, P., Vessman, B. & Mitri, S. Toxicity drives facilitation between 4 bacterial species. Proc. Natl. Acad. Sci. USA 116, 15979–15984 (2019).
    Google Scholar 
    Lofgren, L. et al. Suillus: an emerging model for the study of ectomycorrhizal ecology and evolution. New Phytol. 242, 1448–1475 (2024).
    Google Scholar 
    Voller, F., Ardanuy, A., Taylor, A. F. S. & Johnson, D. Maintenance of host specialisation gradients in ectomycorrhizal symbionts. New Phytol. 242, 1426–1435 (2024).
    Google Scholar 
    Frey-Klett, P., Garbaye, J. & Tarkka, M. The mycorrhiza helper bacteria revisited. New Phytol. 176, 22–36 (2007).
    Google Scholar 
    Tedersoo, L., Bahram, M. & Zobel, M. How mycorrhizal associations drive plant population and community biology. Science 367, eaba1223 (2020).
    Google Scholar 
    Peng, L. et al. A facultative ectomycorrhizal association is triggered by organic nitrogen. Curr. Biol. 32, 5235–5249.e5237 (2022).
    Google Scholar 
    Rasheed, Z. et al. Production of some secondary metabolites of antibiotic nature from mycorrhizal helper bacteria (MHB) associated with conifers. Pol. J. Environ. Stud. 33, 4325–4334 (2024).
    Google Scholar 
    Singh, B. K. & Walker, A. Microbial degradation of organophosphorus compounds. FEMS Microbiol. Rev. 30, 428–471 (2006).
    Google Scholar 
    Yang, F. et al. Genome-wide analysis reveals genetic potential for aromatic compounds biodegradation of Sphingopyxis. Biomed. Res. Int. 2020, 5849123 (2020).
    Google Scholar 
    Zhao, Q. et al. Comparative genomic analysis of 26 Sphingomonas and Sphingobium strains: dissemination of bioremediation capabilities, biodegradation potential and horizontal gene transfer. Sci. Total Environ. 609, 1238–1247 (2017).
    Google Scholar 
    Tong, S. et al. Characterization of a fungal competition factor: production of a conidial cell-wall-associated antifungal peptide. PLoS Pathog. 16, e1008518 (2020).
    Google Scholar 
    Zhong, Z. X. et al. Differential gene expression profiling analysis in Pleurotus ostreatus during interspecific antagonistic interactions with Dichomitus squalens and Trametes versicolor. Fungal Biol. 121, 1025–1036 (2017).
    Google Scholar 
    Mochizuki, K. et al. The ASCT/SCS cycle fuels mitochondrial ATP and acetate production in Trypanosoma brucei. Biochim. Biophys. Acta Bioenerg. 1861, 148283 (2020).
    Google Scholar 
    Labbe, J. L., Weston, D. J., Dunkirk, N., Pelletier, D. A. & Tuskan, G. A. Newly identified helper bacteria stimulate ectomycorrhizal formation in Populus. Front. Plant Sci. 5, 579 (2014).
    Google Scholar 
    Al Farraj, D. A., Alkufeidy, R. M., Alkubaisi, N. A. & Alshammari, M. K. Polynuclear aromatic anthracene biodegradation by psychrophilic Sphingomonas sp., cultivated with Tween-80. Chemosphere 263, 128115 (2021).
    Google Scholar 
    Park, J. et al. Sulfur metabolism-mediated fungal glutathione biosynthesis is essential for oxidative stress resistance and pathogenicity in the plant pathogenic fungus Fusarium graminearum. mBio 15, e0240123 (2024).Hanna, V. S., Abd El-Ghany, M. N., Ibrahim, M. I. M., Abdel-Rahman, T. M. & Tallima, H. Novel approaches to Mortierella alpina identification and arachidonic acid production optimization. ACS Omega 9, 34456–34463 (2024).
    Google Scholar 
    Agerer, R. Exploration types of ectomycorrhizae: a proposal to classify ectomycorrhizal mycelial systems according to their patterns of differentiation and putative ecological importance. Mycorrhiza 11, 107–114 (2001).
    Google Scholar 
    van der Linde, S. et al. Environment and host as large-scale controls of ectomycorrhizal fungi. Nature 558, 243–248 (2018).
    Google Scholar 
    Põlme, S. et al. Biogeography of ectomycorrhizal fungi associated with alders (Alnus spp.) in relation to biotic and abiotic variables at the global scale. New Phytol. 198, 1239–1249 (2013).
    Google Scholar 
    Reis, F. et al. Ectomycorrhizal fungal diversity and community structure associated with cork oak in different landscapes. Mycorrhiza 28, 357–368 (2018).
    Google Scholar 
    Edwards, J. et al. Structure, variation, and assembly of the root-associated microbiomes of rice. Proc. Natl. Acad. Sci. USA 112, E911–E920 (2015).
    Google Scholar 
    He, X. H., Yang, Y. Z. & Yuang, Z. L. Protocol for sampling of root and rhizosphere soils from trees in natural fields. Bio 101, e2003655 (2021).
    Google Scholar 
    Beckers, B., Op De Beeck, M., Weyens, N., Boerjan, W. & Vangronsveld, J. Structural variability and niche differentiation in the rhizosphere and endosphere bacterial microbiome of field-grown poplar trees. Microbiome 5, 25 (2017).
    Google Scholar 
    Chelius, M. K. & Triplett, E. W. The diversity of archaea and bacteria in association with the roots of Zea mays L. Microb. Ecol. 41, 252–263 (2001).
    Google Scholar 
    Bodenhausen, N., Horton, M. W. & Bergelson, J. Bacterial communities associated with the leaves and the roots of Arabidopsis thaliana. PloS ONE 8, e56329 (2013).
    Google Scholar 
    Gardes, M. & Bruns, T. D. ITS primers with enhanced specificity for basidiomycetes-application to the identification of mycorrhizae and rusts. Mol. Ecol. 2, 113–118 (1993).
    Google Scholar 
    White, T. J., Bruns, T., Lee, S. & Taylor, J. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. PCR Protocols. https://doi.org/10.1016/b978-0-12-372180-8.50042-1 (Academic Press, 1990).Põlme, S. et al. FungalTraits: a user-friendly traits database of fungi and fungus-like stramenopiles. Fungal Divers. 105, 1–16 (2021).
    Google Scholar 
    Kim, M., Oh, H. S., Park, S. C. & Chun, J. Towards a taxonomic coherence between average nucleotide identity and 16S rRNA gene sequence similarity for species demarcation of prokaryotes. Int. J. Syst. Evol. Microbiol. 64, 346–351 (2014).
    Google Scholar 
    Deveau, A. et al. The mycorrhiza helper Pseudomonas fluorescens BBc6R8 has a specific priming effect on the growth, morphology and gene expression of the ectomycorrhizal fungus Laccaria bicolor S238N. New Phytol. 175, 743–755 (2007).
    Google Scholar 
    Zhao, L., Wu, X. Q., Ye, J. R., Li, H. & Li, G. E. Isolation and characterization of a mycorrhiza helper bacterium from rhizosphere soils of poplar stands. Biol. Fertil. Soils 50, 593–601 (2014).
    Google Scholar 
    Founoune, H. et al. Interactions between ectomycorrhizal symbiosis and Fluorescent pseudomonads on Acacia holosericea: isolation of mycorrhiza helper bacteria (MHB) from a Soudano-Sahelian soil. FEMS Microbiol. Ecol. 41, 37–46 (2002).
    Google Scholar 
    Zi, H. Y. et al. Co-inoculation with rhizobacterial community and an ectomycorrhizal fungus promotes poplar ectomycorrhization. Appl. Soil Ecol. 206, 105833 (2025).
    Google Scholar 
    Yu, L. et al. Rhizosphere microbiome of forest trees is connected to their resistance to soil-borne pathogens. Plant Soil 479, 143–158 (2022).
    Google Scholar 
    Cavanaugh, J. E. & Neath, A. A. The Akaike information criterion: background, derivation, properties, application, interpretation, and refinements. Wiley Interdiscip. Rev. Comput. Stat. 11, e1460 (2019).Muggeo, V. Segmented: an R package to fit regression models with broken-line relationships. R New. 8, 20–25 (2008).Hu, W. G. et al. Aridity-driven shift in biodiversity-soil multifunctionality relationships. Nat. Commun. 12, 5350 (2021).
    Google Scholar 
    Friedman, J. & Alm, E. J. Inferring correlation networks from genomic survey data. PLoS Comput. Biol. 8, e1002687 (2012).
    Google Scholar 
    Herren, C. M. & McMahon, K. D. Cohesion: a method for quantifying the connectivity of microbial communities. ISME J. 11, 2426–2438 (2017).
    Google Scholar 
    Dufrêne, M. & Legendre, P. Species assemblages and indicator species: the need for a flexible asymmetrical approach. Ecol. Monogr. 67, 345–366 (1997).
    Google Scholar 
    Gweon, H. S. et al. Contrasting community assembly processes structure lotic bacteria metacommunities along the river continuum. Environ. Microbiol. 23, 484–498 (2021).
    Google Scholar 
    Download referencesAcknowledgementsThis work was supported by 32430069, W2412011 (to X.L.); The Jiangsu Special Fund on Technology Innovation of Carbon Dioxide Peaking and Carbon Neutrality BE2022420 (to X.L.). We appreciate Professor Nan Yang of Nanjing Forestry University for providing the Tuber strains.Author informationAuthors and AffiliationsState Key Laboratory of Tree Genetics and Breeding, Nanjing Forestry University, Nanjing, ChinaHaiyun Zi, Zhe Hua, Yun Wang, Yangwenke Liao, Shuikuan Bei, Fuliang Cao & Xiaogang LiSchool of Tea and Coffee, Puer University, Puer, ChinaHaiyun ZiLaboratorio de Biodiversidad y Funcionamiento Ecosistémico, Instituto de Recursos Naturales y Agrobiología de Sevilla (IRNAS), CSIC, Sevilla, SpainManuel Delgado-BaquerizoAuthorsHaiyun ZiView author publicationsSearch author on:PubMed Google ScholarZhe HuaView author publicationsSearch author on:PubMed Google ScholarYun WangView author publicationsSearch author on:PubMed Google ScholarYangwenke LiaoView author publicationsSearch author on:PubMed Google ScholarShuikuan BeiView author publicationsSearch author on:PubMed Google ScholarFuliang CaoView author publicationsSearch author on:PubMed Google ScholarManuel Delgado-BaquerizoView author publicationsSearch author on:PubMed Google ScholarXiaogang LiView author publicationsSearch author on:PubMed Google ScholarContributionsX.L. conceived the project, designed the experiments. H.Z. and Y.W. conducted data curation, methodology, and writing of the original draft. M.D.-B. contributed to data interpretation, writing – review & editing. Z.H., Y.L., S.B., F.C., and M.D.-B. worked on the manuscript. All authors have discussed the results, read and approved the contents of the manuscript.Corresponding authorCorrespondence to
    Xiaogang Li.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 InformationReporting SummaryPeer Review fileRights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleZi, H., Hua, Z., Wang, Y. et al. Mycorrhizal colonization of dryland tree establishment depends on soil microbial cooperation.
    Nat Commun (2025). https://doi.org/10.1038/s41467-025-67797-zDownload citationReceived: 24 September 2025Accepted: 09 December 2025Published: 29 December 2025DOI: https://doi.org/10.1038/s41467-025-67797-zShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Comparative effects of synthetic and natural hydrogels enriched with fertilizer on poppy yield and soil health in drought-prone conditions

    AbstractThe negative effects of agricultural drought are particularly pronounced in spring crops, which are generally less tolerant to dry periods. One such crop frequently affected by drought is poppy (Papaver somniferum L.). Hydrogels enriched with fertilizer represent a promising technology to enhance water availability for plants and improve nutrient uptake from applied fertilizers. The aim of this research was to compare the effects of standard fertilizer (NPKS), a natural-based (NHA) hydrogel, a synthetic hydrogel (SAP), and both hydrogels enriched with fertilizer (NHA-NPKS and SAP-NPKS) on culinary poppy yield, the agronomic efficiency of N fertilization (AEN) and soil microbial activity. Each treatment was applied in two dosages (I and II). Results from a three-year field experiment showed that the application of SAP-NPKS at the lower dose (I) significantly increased seed yield. The highest AEN was also observed in the SAP-NPKS I treatment. The highest seed yield overall was achieved with the higher dose of the natural-based hydrogel enriched with fertilizer (NHA-NPKS II). Furthermore, the use of NHA and NHA-NPKS significantly increased soil microbial activity. These findings suggest that fertilizer-enriched natural-based hydrogels are a promising approach for improving soil moisture retention and nutrient availability, particularly under drought conditions in poppy cultivation.

    Similar content being viewed by others

    Synthesis and application of a multifunctional poly (vinyl pyrrolidone)-based superabsorbent hydrogel for controlled fertilizer release and enhanced water retention in drought-stressed Pisum sativum plants

    Article
    Open access
    12 November 2024

    Improvement of tomato yield and quality using slow release NPK fertilizers prepared by carnauba wax emulsion, starch-based latex and hydrogel nanocomposite combination

    Article
    Open access
    10 July 2023

    Biopolymeric superabsorbent hydrogels enhance crop and water productivity of soybean–wheat system in Indo-Gangetic plains of India

    Article
    Open access
    13 July 2022

    Introduction Climate change is increasingly altering environmental conditions, directly affecting the cultivation of field crops. The rise in temperature and shifts in precipitation patterns have led to a higher incidence of droughts1, which adversely impact crop production and ecosystem services. Yield losses due to drought stress depend on the timing, duration and severity of the drought2. The negative effects of agricultural drought are particularly pronounced in spring-sown crops, which are generally less tolerant to water deficit. One such a crop frequently affected by drought is poppy (Papaver somniferum L.). This oilseed crop is especially vulnerable to unfavourable weather conditions early in the growing season, particularly during germination and emergence3. Efficient soil moisture management is therefore crucial for successful poppy cultivation.The use of synthetic superabsorbent polymers (SAPs) to enhance the water retention capacity of topsoil has been practiced for over two decades4. These polymers can absorb and subsequently release many times their weight in water to support plant growth over time5. As a result, they increase the soil´s water holding capacity6 and help to mitigate plant drought stress7. In addition, hydrogels reduce erosion and nutrients leaching during heavy rainfall, improve soil structure and prevent compaction8. Despite the efforts to use hydrogels based on natural polymers or their organic-inorganic hybrids, yet, the mostly used hydrogels are of synthetic origin (abbreviated here as SAPs), such as acrylate and acrylamide monomers9. The popularity of SAPs is caused particularly due to their low production demands and costs10. The beneficial properties of SAPs are well-documented, but their introduction into soil systems may also pose several risks. A major concern is the persistence of polyacrylic acid due to its extremely low biodegradation rates in soil (e.g., 0.2–0.5% over year)11,12.In accordance to Commission Delegated Regulation (EU) 2024/277013, continuous use of SAPs for water retention improvement in soil is conditional. It is required an ultimate degradation of at least 90% of SAPs (relative to the reference material) within 48 months plus the indicated functionality period. Second option includes mineralization of at least 90% measured by evolved CO2, within the same timeframe (according to the test method EN ISO 17556:201914). Nevertheless, due to the low biodegradability of SAPs, current research efforts turned to developing biodegradable, environment-friendly alternatives15,16,17,18,19.Indeed, over the last decade, natural-based hydrogel alternatives (NHAs) have shown promise as eco- friendly and cost-effective substitutes. A specific group, inorganic hydrogels, appeared to be limited by low swelling capacity and adverse effects on soil fertility20,21. This shifted the attention to NHAs derived from biopolymers such as polysaccharides22 (e.g. cellulose, starch, chitosan or various gums) or proteins7 (e.g. gelatine).In addition to water availability, adequate nutrient supply is critical for optimal plant growth. For poppy, the most important nutrients include nitrogen (N), phosphorus (P), potassium (K), and sulphur (S). Nitrogen is essential for synthesis of amino acids, nucleic acids, enzymes and chlorophyll, playing a key role in biomass production23,24. Phosphorus is involved in synthesis of nucleic acids and phospholipids, respiration, glycolysis, lipid metabolism and energy transfer25. Potassium contributes to ion homeostasis, osmoregulation, enzyme activation, and membrane protein transport26. Sulphur is critical for the synthesis of sulphur-containing amino acids (e.g. cysteine, methionine) and certain vitamins27 and it supports vegetative growth28.These macronutrients are usually supplied through mineral fertilizers. However, a significant portion is often lost through leaching into deeper soil layers, immobilization in soil, volatilization, runoff29, thereby reducing nutrient-use efficiency. As a result, only about 45% of applied nitrogen fertilizer is typically utilized by crops30. Therefore, improving synchronization between nutrient availability and crop demand is crucial for both economic and environmental sustainability.Sustainable agriculture aims to introduce innovative plant nutrition systems that enhance fertilizer efficiency. One such strategy involves the application of NHA-based hydrogels in combination with mineral fertilizers to simultaneously improve soil water retention and nutrient availability. These bio-based polymers, when combined with conventional mineral fertilizers, can potentially hold substantial quantities of water and nutrients, releasing them in sync with plant demand. In case of SAPs their capacity to serve as carriers and regulators of nutrient release, reducing nutrient losses while sustaining plant growth have already been well-documented31,32. In contrast, broader adoption of NHAs is still limited by gaps in understanding the mechanism of nutrient release, the impacts on soil physical, chemical and biological properties as well as on plant root development33. Nonetheless, several studies have already indicated that NHAs can bind nutrients and release them in a controlled manner22. Furthermore, multicomponent NHAs have been found to slow nitrogen release, enhance soil moisture retention, and partially mitigate the environmental risks associated with SAPs18,34.The aim of this study was to evaluate the multi-year effect of SAPs and NHAs enriched with fertilizer (NPKS) on the seed yield of culinary poppy. To the best of our knowledge, the impact of specific nutrient-enriched hydrogels on culinary poppy has not yet been investigated in this context. The main hypothesis was that fertilizer-enriched natural hydrogels would achieve equal or superior yield outcomes compared to synthetic SAPs. To test this hypothesis, a three-year field experiment (2022–2024) was conducted under real agricultural conditions.Materials and methodsExperimental locality and climate-soil conditionsThe effect of fertilizer-enriched natural and synthetic hydrogels on poppy yield was evaluated in a three-year (2022–2024) small-plot field experiment. The trial was conducted at the Žabčice experimental station in South Moravia, Czech Republic (49°1′18.658″ N, 16°36′56.003″ E), at an elevation 184 m above sea level. The site is characterized by mild, wet winters and warm, somewhat dry summers, with an average annual temperature 10.1 °C and annual precipitation of approximately 490 mm. According to the Köppen climate classification, the region falls within the “Cfb” category (temperate oceanic climate). The total precipitation and average air temperature during the experimental growing seasons were 121 mm/11.8 °C (2022), 159 mm/11.0 °C (2023), and 205 mm/14.4 °C (2024). Average monthly temperatures and precipitations during the experimental period, along with the 1991–2020 climatic norm, are presented in Fig. 1.Fig. 1Weather conditions during the field experiment (2022–2024).Full size imageThe experiment was conducted on a single field (240 m × 150 m), which was divided into three experimental Sect. (80 m × 150 m). Each year, poppy was grown on a different section, always following a spring barley pre-crop. Key physicochemical properties of the topsoil (0–30 cm) over the three years are presented in Table 1.Table 1 Physicochemical properties of the experimental soil (0–30 cm depth).Full size tableExperimental design and treatmentsThe objective was to evaluate the effect of two types of hydrogels (synthetic and natural) enriched with nutrients (N, P, K, S) on poppy (Papaver somniferum L.) yield. The natural hydrogel (NHA) was prepared from potato starch (AGRANA Beteiligungs-AG, Konstanz, Germany), glycerol (PENTA, Ltd., Prague, Czechia), clinoptilolite zeolite (particle size 2 mm; Rosteto, Jindrichuv Hradec, Czechia) and potassium polyacrylate (Falconry, Kozmice, Czechia). The final NHA composition consisted of 86wt.% starch- glycerol mixture (43wt.% glycerol), 7 wt% potassium polyacrylate, and 7 wt% zeolite. NHA was prepared by thermoplasticization at 140 °C by passing one cycle in a hot-melt extruder using citric acid as a crosslinking agent. The synthetic superabsorber polymer (SAP) treatment consisted of 100% potassium polyacrylate. Details on NHA preparation and nitrogen release characteristics are described by Skarpa et al.22.The nutrient source was NPKS fertilizer (YARA Mila Complex 12-11-18-8; YARA Agri Czech Republic, Prague, Czechia). The hydrogels, fertilizer and fertilizer-enriched hydrogels were applied in two dosage levels (I and II). The ratio of hydrogel to fertilizer was adjusted to ensure equal application rates of nitrogen (24 and 48 kg/ha) and hydrogel (15 and 30 kg/ha) in both dosage levels (Table 2).Table 2 Treatment design and application rates.Full size tableThe blue-seed poppy variety “MS Harlekyn” (National Agricultural and Food Center, Luzianky, Slovak Republic) was used as a model crop. The experiment followed a Randomized Complete Block Design. Each treatment was replicated 4 times (4 plots per treatment) each year (experimental section). The area of each plot was 15 m2 (10 × 1,5 m). The allocation of replicates across the area of each experimental section was consistent for all years (Figure S1).Fertilizers and hydrogels were manually applied to individual plots one day before sowing and incorporated into the soil immediately after application. Sowing took place on 28 February 2022, 1 March 2023, and 19 March 2024.Harvesting was carried out after physiological maturity (27 July 2022, 20 July 2023, and 22 July 2024) using a Haldrup C-85 plot harvester (Haldrup GmbH, Ilshofen, Germany). Seed yield was measured using a digital scale (KERN DS 60K0.2, KERN and Sohn GmbH, Germany). Grain moisture content was determined with a portable grain moisture meter Wile 78 Crusher (Farmcomp Oy, Tuusula, Finland) and yield was standardized to 8.0% moisture and expressed in tons per hectare (t/ha). The 1000-seeds weight was determined using a scale KERN ARJ 220-4 M, KERN and Sohn GmbH (Balingen, Germany).The agronomic efficiency (AE)37 was expressed for the fertilized treatments as the increase in seed yield (kg) per unit of nitrogen applied (kg) according to the Eq. 1) and per unit of hydrogel applied (kg) according to the Eq. 2):$$:{text{AE}}_{text{N}}text{(kg/kg)=}frac{{text{Y}}_{text{FERT}}text{(t/ha)}:-:{text{Y}}_{text{CONTROL}}text{(t/ha)}}{{text{N}}_{text{DOSE}}text{(kg/ha)}}timestext{1000}$$
    (1)
    $$:{text{AE}}_{text{H}}text{(kg/kg)=}frac{{text{Y}}_{text{FERT}}text{(t/ha)}:-:{text{Y}}_{text{CONTROL}}text{(t/ha)}}{{text{Hydrogel}}_{text{DOSE}}text{(kg/ha)}}times text{1000}$$
    (2)
    where AEN is agronomic efficiency of nitrogen, AEH is agronomic efficiency of hydrogel, YFERT is fertilized treatment yield of poppy seed, YCONTROL is unfertilized treatment yield of poppy seed, NDOSE is nitrogen dose applied by fertilizer, and HydrogelDOSE is hydrogel dose applied.Soil sampling (0–15 cm depth) was carried out after poppy each year. Fresh fine samples were stored at 4 °C and analyzed for dehydrogenase activity (DHA) using 2,3,5-triphenyltetrazolium chloride (TTC) method38, and basal respiration (BR) using MicroResp® device (The James Hutton Institute, Scotland) according to Campbell et al.39.Economic analysisA partial budget analysis40 was performed to assess the cost effectiveness of hydrogels treatments on poppy seed production. This procedure considers only major differences between treatments (fertilization) without considering all costs and benefits. Therefore, only the cost of fertilizer or/and hydrogels, their applications (12 € for each treatment) and price of poppy seed were considered. All non-fertility costs (e.g., seed costs, field operations, plant protection) were held constant across treatments and were not included in the calculation. The cost of 1 ton of used fertilizer (YARA Mila Complex) was 840 €; the cost of 1 ton of SAP and NHA hydrogels was 12 000 and 5 400 € respectively. The prices of fertilizers for treatments were recalculated according to the corresponding applied doses (Table 2). The price of harvested commodity (culinary poppy seed) was 2400 €/t. The prices were based on the actual market values at the end of 2024.However, the inherent volatility input (fertilizer/hydrogels) and output (seed of poppy) prices represents a challenge for accurate economic analysis. This approach focuses on practical economic variability; therefore, scenario-based sensitivity analysis was preferred over statistical confidence intervals, as it more realistically represents uncertainty arising from market and climatic fluctuations. Therefore, additional sensitivity analysis41 was performed under three different scenarios to accommodate possible market dynamics and to assess the effects of input and output price changes on the compared treatments of fertilization. These scenarios were: (Sc. 1) increase cost of hydrogels/fertilizer by 10% but fixed commodity price, (Sc. 2) increase commodity price by 10% with fixed hydrogels/fertilizer cost and (Sc. 3) increase cost of hydrogels/fertilizer by 10% and decrease commodity price by 10% (worst case scenario from farmers’ perspective). The average yield of poppy over the three years of the experiment was used for the economic evaluation. Confidence intervals for average yields were not included, as they would represent within-year experimental variability rather than the economic uncertainty captured by the scenario-based sensitivity analysis, which better reflects market and climatic risks influencing profitability.Statistical analysisThe effect of the hydrogels, fertilizer and their mixtures were assessed using ANOVA. Before performing ANOVA, the assumptions of normality and homogeneity of variances were tested using the Shapiro–Wilk and Levene’s tests, respectively. ANOVA was used to evaluate the effects of hydrogel type, dose, fertilizer, and their combinations. Two model structures were used:

    (i)

    Per-year analyses, performed as a one-way ANOVA with Treatment as a fixed factor and Plot as a random factor (yield, 1000 seed weight, AEN, AEH, DHA, and BR), and.

    (ii)

    Combined analyses, performed as a mixed two-way ANOVA with Year and Treatment as fixed factors and Plot (Year) as a random factor.

    When appropriate, a factorial ANOVA including the main effects of Hydrogel type and Dose, and their interaction (Type × Dose), was used to partition the total variability. For each model, the F-statistic, degrees of freedom (df), and p-values for main effects and their interactions were calculated and are reported in Supplementary Tables S1–S3. The effect sizes were expressed as eta-squared (η² = SSeffect/SStotal) and partial eta-squared (partial η² = SSeffect/(SSeffect + SSerror)), representing the proportion of total variance explained by each factor. After a significant omnibus F-test (p ≤ 0.05), Fisher’s Least Significant Difference (LSD) test was applied for post-hoc multiple comparisons among treatment means. All statistical analyses were conducted using Statistica 14 CZ software42. Results are expressed as means ± standard deviations (SD) or standard errors (SE), as appropriate.ResultsSeed yield of poppy and agronomic efficiencyThe effects of hydrogels (SAP, NHA) enriched and not-enriched with fertilizer applied at two doses (I and II) on poppy seed yield are presented in Fig. 2. In all years, it is evident that the higher nutrient dose (II) applied in conventional fertilizer (NPKS) relatively increased seed yield in comparison with the lower dose (NPKS I) by 1.6% (2023), 4.7% (2022) and 5% (2024). The increase in yield of poppy seed caused by the higher NPKS dose was significant compared to the control in 2023 and 2024.The yield response was also influenced by hydrogel dose, which accounted for 32.3% (η² = 0.323, Partial η² = 0.487, p = 0.335) of total seed yield variability, while hydrogel type explained 7.6% (η² = 0.076, Partial η² = 0.183, p = 0.638). A higher dose of synthetic SAP relatively reduced seed production (by 5.3% on average), whereas a higher dose of a natural-based hydrogel increased poppy seed yield: Control (1.15 t/ha, 100%) ˂ NHA I (1.22 t/ha, 106.5%) ˂ NHA II (1.29 t/ha, 112.3%).The highest seed yields were observed in all years with fertilizer-enriched hydrogels. At the lower fertilizer rate (I), except in 2022, the combination of fertilizer with synthetic SAP had a higher effect on production compared to NHA (Fig. 2). This corresponded to the effect of pure SAP. At the higher nutrient rate (II), a higher increase in poppy yield for the fertilizer-NHA combination was observed (in 2023 and 2024). The effect of soil application of the tested fertilizers on seed yield, expressed as a mean for lower dose of hydrogels/fertilizer (I), was as follows: 1.15 ± 0.12 t/ha (Control) ˂ 1.22 ± 0.14 t/ha (NHA) ˂ 1.27 ± 0.11 t/ha (NPKS) ˂ 1.29 ± 0.15 t/ha (NHA-NPKS) ˂ 1.32 ± 0.18 t/ha (SAP) ˂ 1.36 ± 0.22 t/ha (SAP-NPKS). In contrast, when higher rates (II) were used, the effect of the tested fertilizer types was as follows: 1.15 ± 0.12 t/ha (Control) ˂ 1.26 ± 0.17 t/ha (SAP) ˂ 1.29 ± 0.17 t/ha (NHA) ˂ 1.32 ± 0.10 t/ha (NPKS) ˂ 1.36 ± 0.13 t/ha (SAP-NPKS) ˂ 1.38 ± 0.18 t/ha (NHA-NPKS).Fig. 2Effects of fertilization on poppy seed yield (t/ha) in 2022, 2023, 2024 (a), and average of three years (b). Control: treatment without fertilization; NHA: fertilized with bio- natural-based hydrogel; SAP: fertilized with synthetic hydrogel; NPKS: fertilized with NPKS fertilizer; NHA-NPKS: fertilized with NPKS fertilizer-enriched bio- natural-based hydrogel; SAP-NPKS: fertilized with NPKS fertilizer-enriched synthetic hydrogels. Roman numerals I and II indicate hydrogels/fertilizer rates. The columns marked by different lower-case letters indicate significant differences among treatments (each year was evaluated separately). The columns represent the arithmetic means (n = 4), standard deviation is expressed by error bars. The values F, df, and P for main effects and interactions are given in Tables S1.Full size imageThe 1000-seed weight was not significantly affected by fertilization in 2022 and 2023 (Table 3). In 2024, the significantly highest poppy seed weight was found in the higher fertilizer rate (II) treatments as follows: NHA-NPKS ˂ NPKS ˂ SAP-NPKS. Consistent with the results of the 3rd year of testing, the relatively highest average seed weight was found on the treatments fertilized with higher rates of pure fertilizer and hydrogels enriched by fertilizer. Their increased doses (II) resulted in an increase of poppy seed weight compared to the lower doses (I), by 6.6% (NPKS), 7.2% (NHA-NPKS) and 11.4% (SAP-NPKS), respectively. Pure hydrogels did not affect seed weight significantly.Table 3 Effects of fertilization on 1000 seed weight (g).Full size tableThe agronomic efficiency of nitrogen (AEN) and hydrogel (AEH) is shown in Table 4.The lower rate of nitrogen applied by NPKS fertilizer resulted in significantly higher AEN compared to the higher rate in the average of three years (Table 4). The largest difference in AEN between nitrogen rates was observed in 2023, where 1 kg of nitrogen applied at the lower rate increased poppy seed yield by 6.9 kg, while the yield increase at the higher rate was 3.8 kg of seed.The agronomic nitrogen efficiency of fertilizer-enriched hydrogels applied at a lower rate was significantly higher when using synthetic SAP. Significant increase in AEN for SAP-NPKS I compared to NPKS I was found in 2023 (+ 46.4%) and 2024 (+ 160.7%), averaging 69.2% over the three years (Table 4). An average increase in AEN (+ 15.4%) was also observed with the use of fertilizer-enriched NHA (NHA-NPKS I), but not significant. Higher rates of hydrogels enriched by fertilizer did not statistically affect the efficiency of applied nitrogen. The relatively highest AEN value was found for the NHA-NPKS II treatment (4.9; 100%), followed by SAP-NPKS II (91.8%) and NPKS II (71.4%). The total variability of AEN was significantly influenced mainly by the dose of hydrogel (η² = 0.570, Partial η² = 0.851, p = 0.021) while the type of hydrogel explained 11.7% (η² = 0.117, Partial η² = 0.539, p = 0.286).With pure hydrogel applied at a lower rate, poppy seed yield increased significantly with synthetic SAP. The average AEH for SAP I was 11.7 (i.e. the seed yield increased by 11.7 kg due to the application of 1 kg of SAP). The agronomic efficiency of the synthetic hydrogel was more than twice as high compared to NHA (Table 4). In contrast, for the higher dose of hydrogel, the efficiency of NHA was similar to the effect of its lower dose, whereas in the case of SAP, AEH was significantly reduced (more than threefold decrease).The effect of hydrogels (AEH) on poppy yield increased when used in combination with fertilizers. In the case of natural hydrogel enriched by fertilizer (NHA-NPKS) compared to its pure form (NHA), a significant increase in AEH was observed for both doses (I + 92%, II + 68.1%). In the case of synthetic hydrogel, an increase in AEH was also observed between the SAP-NPKS and SAP, but significantly only for the higher dose (+ 89.5%).Table 4 Effects of fertilization on agronomic efficiency of nitrogen (AEN), and agronomic efficiency of hydrogel (AEH).Full size tableSoil microbial activity and biomassNo significant effects of any treatment of hydrogels (SAP, NHA) applied either solely or with NPS was observed on soil dehydrogenase activity (DHA) in 2022 (Table 5). In the next year 2023, all types of amendments (except of NHA I) increased DHA in comparison to Control value, and increased value in SAP I, which was even higher in combination with higher NPKS (SAP-NPKS II) as well as in all treatments with higher or/and combined NHA amendment (NHA-NPKS I, NHA II, NHA-NPKS II). Moreover, NHA II enhanced soil DHA significantly more than both doses of NPKS (I and II), showing the highest enzyme values in 2023, 2024 and in 3-year average (Table 5). In 2024, only treatments with sole NHA (I, II), NHA-NPKS I, and NPKS I were increased over Control, while SAP II was decreased. In 3-year average, SAP applied solely (in low dose I) or combined (I, II) increased DHA over Control values but not compared to NPKS (I, II). Only NHA-NPKS I and NHA II enhanced DHA more than amendment of fertilizers.Table 5 Effects of fertilization on dehydrogenase activity (DHA) of microbial biomass and basal soil respiration (BR).Full size tableIn 2022, soil basal respiration (BR) was decreased compared to Control in treatments with high fertilizer dose (NPKS II), applied solely or combined with NHA (Table 5). In 2023, only NPKS I (low dose) exerted negative effect on BR. In 2024, NPKS II again solely or combined (SAP-NPKS II) showed significant decrease in BR, while NPKS I and NHA II increased the values over Control. In 3-year average, mainly insignificantly different or negative effects of tested amendments on BR were found, the strongest BR reduction was derived by NPKS II, SAP I, NHA-NPKS I and II (both; Table 5).Economic evaluation of poppy productionTable 6 describes the results of the economic evaluation of poppy production for the tested fertilization treatments under different price scenarios, considering the variable price of fertilizers and commodity (poppy seed). At lower fertilizer doses (I), the application of pure SAP (SAP I) and SAP enriched by fertilizer (SAP-NPKS I) was the most profitable in each scenario. The net profit of the NHA I treatment was approximately 38% of the SAP I profit. In the case of fertilizer-enriched hydrogels, the difference between net profit of SAP and NHA was lower (Table 6).Higher doses of hydrogels and their fertilizer-enriched forms increased the cost of poppy production and reduced profits. The use of higher doses of SAP (SAP II, SAP-NPKS II) was unprofitable (loss-making in all scenarios). The higher poppy yield obtained after applying natural hydrogels at a higher dose (II) increased the profit, in the case of using pure NHA in the range of 115–199 €/ha, when using NHA-NPKS depending on the type of scenario, as shown in Table 6 (highest profit in Sc. 2, conversely, loss in the case of Sc. 3).Table 6 Economic evaluation of poppy fertilization.Full size tableDiscussionThe effect of the tested types of hydrogels on poppy (Papaver somniferum L.) seed yield varied depending on the type and application rate. At a lower dose, the synthetic SAP increased seed yield more effectively compared to the natural hydrogel (NHA), whereas the opposite trend was observed at the higher dose. This pattern was consistent for both pure hydrogels and their fertilizer-enriched formulations.Specifically, the lower application rate of pure SAP and SAP combined with fertilizer (SAP-NPKS I) significantly improved poppy seed yield. These results align with prior studies that demonstrated increased crop yields following SAP application due to their high water absorption and retention capacity43, as well as improvements in soil physical and chemical properties44,45. Additionally, SAPs have been shown to reduce nutrient losses46,47, and enhance fertilizer use efficiency45. In the present study, the agronomic efficiency of nitrogen (AEN) was significantly higher in the SAP-NPKS treatment compared to NPKS alone (Table 4).Yield differences following hydrogel application are attributed to the feedstock’s composition, hydrogel formulation and method of application48. Synthetic SAP generally show stronger positive effects on yield than natural hydrogels9. Zheng et al.9 reported an average 12.8% in crop yield from SAP application (95% confidence intervals: 12.1–13.4%, p < 0.01). In our trial, pure SAP application increased poppy yield by 12.4% o average across both doses, whereas pure NHA increased yield by 9.4%. Zheng et al.9 also found a 15.2% increase in oilseed-yields, including poppy, after SAP application, although data specific to poppy remain scarce. One of the few relevant studies showed improved oilseed rape yield under drought and irrigation conditions with the application of anionic cellulose-based polyacrylate hydrogel at 40 kg/ha49.However, high doses of synthetic SAPs may negatively affect plant growth. Situ et al.34 found that excessive use of synthetic SAP reduced biomass and roots and stems–leaves, likely due to ion imbalances (elevated K+ and Na+, reduced Ca2+ and Mg2+). This may explain observed reductions in germination rate, plant height and yield. In our study the SAP II treatment led to a relative yield decline in two of the three trial years. On average, the SAP II treatment yielded 5.6% less than the NHA II treatment.The effect of hydrogels on seed weight and poppy production also depended on weather conditions during vegetation periods. In particular, in relatively dry and cool years (2022 and 2023) seed weight was not significantly affected by fertilization (Table 3). On the contrary, 2024 growing period was characteristic of sufficient water supply and higher temperatures, which supported soil microbial activity (as reflected in DHA and BR, Table 5), in particular in the presence of bio-SAP-based fertilizers. The elevated microbial activity increases soil organic matter turnover and contribute to soil fertility50, with a direct impact on crop yields51,52, as observed in this work.However, at higher SAP rates the reduced growth was observed, which may stem (apart from ion imbalance), from the presence of acrylic acid. Chen et al.53 reported damage to the organizational structure and cellular morphology of maize roots and the membrane system of root cells. Puoci et al.54 described acrylate hydrogel intermediates as cytotoxic. Additionally, excessive SAP can compete with plants for water under drought conditions, potentially exacerbating water stress in plants and thereby reducing yields55,56. This aligns with Zheng et al.9 who concluded that SAP rates > 90 kg/ha do not significantly improve yields and may even be detrimental, while the application rate 45–90 kg/ha exhibited positive results.Although recent research focuses increasingly on natural polymers or organic-inorganic hybrid compounds, most studies still involve synthetic SAPs (acrylate and acrylamide-based)9. Nevertheless, several studies confirm yield benefits from bio-based hydrogels22,57,58,59,60, which offer key advances such as biocompatibility, non-toxicity, and biodegradability. However, “biodegradability” is a broad term and does not necessarily reflect degradation rates. The NHA used in this study, composed primarily of potato starch (86 wt%), is highly biodegradable. Therefore, NHA was superior in enhancing microbial activity as starch degradation products provide carbon source for microorganisms. Based on CO2 released during analysis, Guo et al.61 reported that 78.34% of starch degraded within 14 days. The starch decomposition rate reported in laboratory tests is difficult to achieve under field conditions. Starch and starch based polyurethane materials (starch-polyhydroxyurethanes) lost about 44.1 and 66.4% of their weight, respectively, after 60 days burial in soil62. The biodegradability of NHA was also supported by elevated dehydrogenase activity and basal respiration measured in soil after harvest (Table 5). Soil microbial activity was more strongly stimulated by NHA and NHA-NPKS than by NPKS alone, likely due to excellent biodegradability of starch63 and glycerol64.While SAP also stimulated soil microbiome which is reflected in elevated DHA65, the stimulation was weaker, likely due to slow biodegradation of polyacrylic gels that are primarily degraded fungi such as Phanerochaete chrysosporium66. Although NHA-NPKS enhanced DHA, it did not significantly increase basal respiration over the three-year average. This is likely due to the limited stimulator effect of NPKS amendment, in agreement with study of Kulachkova et al.67, who reported minimal BR after Nitroammofoska-1 application (NPKS 21-10-10-2, similar to the composition of YARA Mila Complex 12-11-18-8) to urban lawn soil.On average, BR was higher following NHA application compared to SAP, by 4.6% at the lower dose, and 5.0% at the higher dose. The poor biodegradability of synthetic SAPs raises environmental concerns68. Their degradation rate is typically 0.45 to 0.82% over 24 weeks depending on soil type but not on soil temperature. Detailed study showed that the polyacrylate superabsorbent main chain degraded in the soils at rates of 0.12–0.24% per 6 months11. Aging via chemical, photolytic, and mechanical stress can lead to fragmentation and formation of microplastic particles, which may leach into deeper soil layers or into adjacent ecosystems, potentially impacting microbial communities and plant growth69,70.Dehydrogenase activity is a well-established indicator of overall microbial activity71. Significantly higher DHA values were observed found in two of the three years and on average in the NHA II treatment (Table 4). Soil treated with NHA and NHA-NPKS exhibited the highest DHA (7.1 a ± 1.8) significantly greater than SAP treatments (6.8 b ± 1.3) and non-hydrogel controls (6.6 c ± 1.5) (p˂0.05). Excessive doses of SAP (based on polyacrylic acid) have been reported to supress microbial respiration in sandy soils72. Soil microbial biomass plays a crucial role in nutrient cycling and natural based hydrogel may further support microbial growth by providing degradable organic substances73, enhancing microbial diversity74, and ultimately improving soil vitality, plant growth and survival rates73.In addition to agronomic and environmental performance, economic viability is crucial for hydrogels adoption. Although economic analysis of hydrogel use remain limited, they are essential to evaluate practical constrains and inform farmers. Commercial synthetic SAPs based on polyacrylic acid are costly despite their high swelling capacity68. Natural hydrogels represent a lower-cost, faster-degrading alternatives with promising market potential48.In this work, a lower dose of potassium polyacrylate (SAP I) resulted in the highest net profit up to 269 €/ha (Sc. 2, Table 5) due to the increase in poppy seed yield. The fertilizer-enriched SAP (SAP-NPKS I) was profitable compared to fertilizer (NPKS I) (average of all calculated scenarios: +20.8 €/ha). The increase in yield and net profit at a lower dose of NHA (NHA I) was also economically beneficial (61–105 €/ha). The use of NHA enriched with fertilizer (NHA-NPKS I) did not exhibit an increased profit compared to SAP. In this context, however, it is also important to consider the environmental compatibility of natural-based hydrogels, even though they may be less economically attractive.These results suggest that poppy is among the crops for which hydrogel application can be economically justified. Yet, profitability depends on crop type. For example, despite yield increases, SAP costs were not offset by revenues in grain crops (net loss of 11 €/ha)9. On the other hand, net profit gains have been documented in maize75, sugarcane76, potatoes44, Indian mustard77, and summer pearl millet78. In our study high-dose NHA (NHA II) indicate an net profit increased by 115 to 199 €/ha, while high-dose SAP (SAP II) appeared to be economically unviable.ConclusionThis study demonstrates that the application of hydrogels, particularly when enriched with fertilizers, can significantly enhance the yield and nutrient-use efficiency of culinary poppy cultivated under drought-prone conditions. While low-dose synthetic SAP treatments provided the highest net economic returns, high-dose SAP applications proved less effective and potentially detrimental due to reduced biodegradability and possible phytotoxicity. In contrast, natural-based hydrogels (NHA), especially when combined with fertilizer, promoted soil microbial activity and showed consistent yield benefits at both application rates. Although the economic return from NHA was generally lower than from SAP, its environmental advantages, such as enhanced biodegradability and stimulation of beneficial soil microbiota, make it a compelling alternative for sustainable agriculture.Overall, natural starch-based hydrogels enriched with fertilizer represent a viable, environmentally friendly strategy for improving soil water retention, nutrient efficiency, and crop performance in poppy cultivation. However, the composition of hydrogels (source of nutrients, e.g., potassium), site-specific conditions such as soil type, climate, and crop response variability must be considered when selecting the appropriate hydrogel type and dose for field application.

    Data availability

    The datasets generated and/or analysed during the current study are available from the corresponding author on reasonable request.
    AbbreviationsNHA:
    natural-based hydrogel
    SAP:
    synthetic hydrogel
    NPKS:
    mineral fertilizer YARA Mila Complex
    NHA-NPKS:
    natural-based hydrogel enriched with fertilizer
    SAP-NPKS:
    synthetic hydrogel enriched with fertilizer
    AEN
    :
    agronomic efficiency of nitrogen fertilization
    AEH
    :
    agronomic efficiency of hydrogel
    DHA:
    dehydrogenase activity
    BR:
    basal respiration
    ReferencesFurtak, K. & Wolińska, A. The impact of extreme weather events as a consequence of climate change on the soil moisture and on the quality of the soil environment and agriculture – A review. Catena (Amst). 231, 107378 (2023).Article 

    Google Scholar 
    Pandey, R. K., Maranville, J. W. & Admou, A. Tropical wheat response to irrigation and nitrogen in a Sahelian environment. I. Grain yield, yield components and water use efficiency. Eur. J. Agron. 15, 93–105 (2001).Article 

    Google Scholar 
    Kundrátová, K. et al. Transcriptomic and proteomic analysis of drought stress response in opium poppy plants during the first week of germination. Plants 10, 1878 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Behera, S. & Mahanwar, P. A. Superabsorbent polymers in agriculture and other applications: a review. Polymer-Plastics Technol. Mater. 59, 341–356 (2020).Article 

    Google Scholar 
    Ghobashy, M. M. et al. Synthesis and application of a multifunctional Poly (vinyl pyrrolidone)-based superabsorbent hydrogel for controlled fertilizer release and enhanced water retention in drought-stressed Pisum sativum plants. Sci. Rep. 14, 27734 (2024).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Yu, J. et al. Superabsorbent polymer properties and concentration effects on water retention under drying conditions. Soil Sci. Soc. Am. J. 81, 889–901 (2017).Article 
    ADS 

    Google Scholar 
    Sarmah, D. & Karak, N. Biodegradable superabsorbent hydrogel for water holding in soil and controlled-release fertilizer. J Appl. Polym. Sci 137, 48495 (2020).Cao, Y., Wang, B., Guo, H., Xiao, H. & Wei, T. The effect of super absorbent polymers on soil and water conservation on the terraces of the loess plateau. Ecol. Eng. 102, 270–279 (2017).Article 

    Google Scholar 
    Zheng, H. et al. Effects of super absorbent polymer on crop yield, water productivity and soil properties: A global meta-analysis. Agric. Water Manag. 282, 108290 (2023).Article 

    Google Scholar 
    Yang, Y. et al. Research advances in superabsorbent polymers. Polym. (Basel). 16, 501 (2024).Article 

    Google Scholar 
    Wilske, B. et al. Biodegradability of a polyacrylate superabsorbent in agricultural soil. Environ. Sci. Pollut. Res. 21, 9453–9460 (2014).Article 

    Google Scholar 
    Liang, D. et al. Degradation of polyacrylate in the outdoor agricultural soil measured by FTIR-PAS and LIBS. Polym. (Basel). 10, 1296 (2018).Article 
    ADS 

    Google Scholar 
    European Union Delegated Regulation (EU) 2024/2770 of the Commission of July 15 2024. Amending Regulation (EU) 2019/1009 of the European Parliament and of the Council with Respect to the Biodegradability Criteria Applicable to Coating Agents and Water Retention Polymers; European Union: Brussels, Belgium, 2024. Preprint at (2024).ISO 17556 Plastics—determination of. the ultimate aerobic biodegradability of plastic materials in soil by measuring the oxygen demand in a respirometer or the amount of carbon dioxide evolved. ISO https://www.iso.org/cms/render/live/en/sites/isoorg/contents/data/standard/07/49/74993.html Preprint at (2019). Dodangeh, F., Nabipour, H., Rohani, S. & Xu, C. Applications, challenges and prospects of superabsorbent polymers based on cellulose derived from lignocellulosic biomass. Bioresour Technol. 408, 131204 (2024).Article 
    PubMed 

    Google Scholar 
    Firmanda, A. et al. Biopolymer-based slow/controlled-release fertilizer (SRF/CRF): nutrient release mechanism and agricultural sustainability. J. Environ. Chem. Eng. 12, 112177 (2024).Article 

    Google Scholar 
    Xiao, X. et al. One-step method to prepare starch-based superabsorbent polymer for slow release of fertilizer. Chem. Eng. J. 309, 607–616 (2017).Article 

    Google Scholar 
    Ribeiro, A. B. et al. Bio-based superabsorbent hydrogels for nutrient release. J. Environ. Chem. Eng. 12, 112031 (2024).Article 

    Google Scholar 
    Wen, P. et al. Rapid synthesis of a corncob-based semi-interpenetrating polymer network slow-release nitrogen fertilizer by microwave irradiation to control water and nutrient losses. Arab. J. Chem. 10, 922–934 (2017).Article 
    ADS 

    Google Scholar 
    Chen, M. et al. Kaolin-Enhanced superabsorbent composites: Synthesis, characterization and swelling behaviors. Polym. (Basel). 13, 1204 (2021).Article 

    Google Scholar 
    Noh, Y. D., Komarneni, S. & Park, M. Mineral-Based slow release fertilizers: A review. Korean J. Soil Sci. Fertil. 48, 1–7 (2015).Article 

    Google Scholar 
    Skarpa, P. et al. Effect of fertilizers enriched with bio-based carriers on selected growth parameters, grain yield and grain quality of maize (Zea Mays L). Eur. J. Agron. 143, 126714 (2023).Article 

    Google Scholar 
    Wang, Q., Li, S., Li, J. & Huang, D. The utilization and roles of nitrogen in plants. Forests 15, 1191 (2024).Article 

    Google Scholar 
    Mohammed, Y. A. et al. Nitrogen fertilizer management for improved grain quality and yield in winter wheat in Oklahoma. J. Plant. Nutr. 36, 749–761 (2013).Article 

    Google Scholar 
    Lovio-Fragoso, J. P. et al. Biochemical and molecular aspects of phosphorus limitation in diatoms and their relationship with biomolecule accumulation. Biology (Basel). 10, 565 (2021).PubMed 

    Google Scholar 
    Johnson, R. et al. Potassium in plants: growth regulation, signaling, and environmental stress tolerance. Plant Physiol. Biochem. 172, 56–69 (2022).Article 
    PubMed 

    Google Scholar 
    Antošovský, J., Škarpa, P., Ryant, P. & Brtnický, M. Waste sulfur from biogas desulphurization: a supplement of Brassica Napus L. nutrition. J. Plant. Nutr. 47, 296–313 (2024).Article 

    Google Scholar 
    Narayan, O. P., Kumar, P., Yadav, B., Dua, M. & Johri, A. K. Sulfur nutrition and its role in plant growth and development. Plant. Signal. Behav. https://doi.org/10.1080/15592324.2022.2030082 (2022).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Lehmann, J. & Schroth, G. Nutrient leaching. In Trees, Crops and Soil Fertility: Concepts and Research Methods 151–166 (CABI Publishing, 2002). https://doi.org/10.1079/9780851995939.0151.Chapter 

    Google Scholar 
    Yu, X., Keitel, C., Zhang, Y., Wangeci, A. N. & Dijkstra, F. A. Global meta-analysis of nitrogen fertilizer use efficiency in rice, wheat and maize. Agric. Ecosyst. Environ. 338, 108089 (2022).Article 

    Google Scholar 
    Islam, M. R. et al. Effects of water-saving superabsorbent polymer on antioxidant enzyme activities and lipid peroxidation in corn (Zea Mays L.) under drought stress. J. Sci. Food Agric. 91, 813–819 (2011).Article 
    PubMed 

    Google Scholar 
    Liu, F. et al. Effects of super-absorbent polymer on dry matter accumulation and nutrient uptake of Pinus pinaster container seedlings. J. For. Res. 18, 220–227 (2013).Article 

    Google Scholar 
    Liao, R., Yang, P., Wu, W. & Ren, S. An inverse method to estimate the root water uptake Source-Sink term in soil water transport equation under the effect of superabsorbent polymer. PLoS One. 11, e0159936 (2016).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Situ, Y. et al. Effects of several superabsorbent polymers on soil exchangeable cations and crop growth. Environ. Technol. Innov. 30, 103126 (2023).Article 

    Google Scholar 
    Gee, G. W. & Bauder, J. W. Particle-size analysis. in 383–411 https://doi.org/10.2136/sssabookser5.1.2ed.c15 (2018).Zbíral, J., Malý, S. & Váňa, M. Soil Analysis 3rd Edn (Central Institute for Supervising and Testing in Agriculture, 2011).Antošovský, J., Škarpa, P. & Ryant, P. The effect of nitrification inhibitor on the yield and quality of triticum aestivum L. and brassica Napus L. – A long-term experiment. Field Crops Res. 328, 109906 (2025).Article 

    Google Scholar 
    Dick, R. P., Breakwell, D. P. & Turco, R. F. Soil enzyme activities and biodiversity measurements as integrative Microbiological indicators. in 247–271 (2015) https://doi.org/10.2136/sssaspecpub49.c15Campbell, C. D., Chapman, S. J., Cameron, C. M., Davidson, M. S. & Potts, J. M. A rapid microtiter plate method to measure carbon dioxide evolved from carbon substrate amendments so as to determine the physiological profiles of soil microbial communities by using whole soil. Appl. Environ. Microbiol. 69, 3593–3599 (2003).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    CIMMYT Economics Program. From Agronomic Data To Farmer Recommendations: an Economics Training Manual (CIMMYT Economics Program, 1988).Mohammed, Y. A., Gesch, R. W., Johnson, J. M. F. & Wagner, S. W. Agronomic and economic evaluations of N fertilization in maize under recent market dynamics. Nitrogen 3, 514–527 (2022).Article 

    Google Scholar 
    StatSoft, I. STATISTICA (Data Analysis Software System) Preprint at (2022).Feng, W. et al. Effects of Polyacrylamide-Based super absorbent polymer and corn straw Biochar on the arid and Semi-Arid salinized soil. Agriculture 10, 519 (2020).Article 

    Google Scholar 
    Hou, X. et al. Superabsorbent polymers influence soil physical properties and increase potato tuber yield in a dry-farming region. J. Soils Sediments. 18, 816–826 (2018).Article 

    Google Scholar 
    Egrinya Eneji, A., Islam, R., An, P. & Amalu, U. C. Nitrate retention and physiological adjustment of maize to soil amendment with superabsorbent polymers. J. Clean. Prod. 52, 474–480 (2013).Article 

    Google Scholar 
    Mikkelsen, R. L. Using hydrophilic polymers to control nutrient release. Fertilizer Res. 38, 53–59 (1994).Article 

    Google Scholar 
    Zheng, T., Liang, Y., Ye, S. & He, Z. Superabsorbent hydrogels as carriers for the controlled-release of urea: experiments and a mathematical model describing the release rate. Biosyst Eng. 102, 44–50 (2009).Article 

    Google Scholar 
    Grabowska-Polanowska, B., Garbowski, T., Bar-Michalczyk, D. & Kowalczyk, A. The benefits of synthetic or natural hydrogels application in agriculture: an overview Article. J. Water Land. Dev. 208–224 https://doi.org/10.24425/jwld.2021.139032 (2022).Badr, E. A. et al. Enhancing Canola yield and photosynthesis under water stress with hydrogel polymers. Phyton (B Aires). 93, 1623–1645 (2024).Article 

    Google Scholar 
    Lin, J. et al. Effect of degradable microplastics, Biochar and their coexistence on soil organic matter decomposition: A critical review. TRAC Trends Anal. Chem. 183, 118082 (2025).Article 

    Google Scholar 
    Shu, X. et al. Organic amendments enhance soil microbial diversity, microbial functionality and crop yields: A meta-analysis. Sci. Total Environ. 829, 154627 (2022).Article 
    PubMed 

    Google Scholar 
    Tautges, N. E., Sullivan, T. S., Reardon, C. L. & Burke, I. C. Soil microbial diversity and activity linked to crop yield and quality in a dryland organic wheat production system. Appl. Soil. Ecol. 108, 258–268 (2016).Article 

    Google Scholar 
    Chen, X., Huang, L., Mao, X., Liao, Z. & He, Z. A comparative study of the cellular microscopic characteristics and mechanisms of maize seedling damage from superabsorbent polymers. Pedosphere 27, 274–282 (2017).Article 

    Google Scholar 
    Puoci, F. et al. Polymer in agriculture: a review. Am. J. Agric. Biol. Sci. 3, 299–314 (2008).Article 

    Google Scholar 
    Shahram, S. & Felora, R. Investigation of superabsorbent polymer and water stress on physiological indexes of maize. J. Adv. Biol. 4, 455–460 (2014).
    Google Scholar 
    Li, H. Y., Zhang, R. & Wang, F. X. Effects of water-retaining agent on soil water movement and water use efficiency of maize sowed with absolved water-storing irrigation. Trans. Chin. Soc. Agricultural Eng. 27, 37–42 (2011).
    Google Scholar 
    Iftime, M. M., Ailiesei, G. L., Ungureanu, E. & Marin, L. Designing Chitosan based eco-friendly multifunctional soil conditioner systems with Urea controlled release and water retention. Carbohydr. Polym. 223, 115040 (2019).Article 
    PubMed 

    Google Scholar 
    Mohamady Ghobashy, M. The application of natural polymer-based hydrogels for agriculture. in Hydrogels Based on Natural Polymers 329–356 (Elsevier, doi:https://doi.org/10.1016/B978-0-12-816421-1.00013-6 (2020).López-Velázquez, J. C. et al. Gelatin–chitosan–PVA hydrogels and their application in agriculture. J. Chem. Technol. Biotechnol. 94, 3495–3504 (2019).Article 

    Google Scholar 
    Supare, K. & Mahanwar, P. A. Starch-derived superabsorbent polymers in agriculture applications: an overview. Polym. Bull. 79, 5795–5824 (2022).Article 

    Google Scholar 
    Guo, W. et al. Introduction of environmentally degradable parameters to evaluate the biodegradability of biodegradable polymers. PLoS One. 7, e38341 (2012).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Ghasemlou, M. et al. Biodegradation of novel bioplastics made of starch, polyhydroxyurethanes and cellulose nanocrystals in soil environment. Sci. Total Environ. 815, 152684 (2022).Article 
    PubMed 

    Google Scholar 
    Cui, C. et al. Recent advances in the preparation, characterization, and food application of starch-based hydrogels. Carbohydr. Polym. 291, 119624 (2022).Article 
    PubMed 

    Google Scholar 
    Pathak, V. Effect of Starch-based Hydrogel on Early Growth of Corn (School of Agricultural & Biological Engineering, 2018).Khodadadi Dehkordi, D. Effects of hydrophilic polymers on soil Water, wheat plant and microorganisms. Appl. Ecol. Environ. Res. 16, 1711–1724 (2018).Article 

    Google Scholar 
    Stahl, J. D., Cameron, M. D., Haselbach, J. & Aust, S. D. Biodegradation of superabsorbent polymers in soil. Environ. Sci. Pollut. Res. 7, 83–88 (2000).Article 

    Google Scholar 
    Kulachkova, S. A., Derevenets, E. N., Korolev, P. S. & Pronina, V. V. The effect of mineral fertilizers on soil respiration in urban lawns. Mosc. Univ. Soil. Sci. Bull. 78, 281–291 (2023).Article 

    Google Scholar 
    Ha, J. et al. Direct measurement of crosslinked surface layer in superabsorbent poly(acrylic acid). Mater. Lett. 228, 33–36 (2018).Article 
    ADS 

    Google Scholar 
    Sojka, R. E., Bjorneberg, D. L., Entry, J. A., Lentz, R. D. & Orts, W. J. Polyacrylamide in agriculture and environmental land management. in 75–162 https://doi.org/10.1016/S0065-2113(04)92002-0 (2007).Ren, X., Wang, L., Tang, J., Sun, H. & Giesy, J. P. Combined effects of degradable film fragments and micro/nanoplastics on growth of wheat seedling and rhizosphere microbes. Environ. Pollut. 294, 118516 (2022).Article 
    PubMed 

    Google Scholar 
    Wolinska, A. & Stepniewsk, Z. Dehydrogenase Activity in the Soil Environment. in DehydrogenasesInTech, https://doi.org/10.5772/48294 (2012).Buchmann, C., Rudolph, S., Neff, J. & Steinmetz, Z. Impact of polyacrylic acid as soil amendment on soil microbial activity under different moisture regimes. Sci. Rep. 15, 19422 (2025).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bana, R. S. et al. Enhanced Pearl millet yield stability, water use efficiency and soil microbial activity using superabsorbent polymers and crop residue recycling across diverse ecologies. Eur. J. Agron. 148, 126876 (2023).Article 

    Google Scholar 
    Li, X., He, J. Z., Hughes, J. M., Liu, Y. R. & Zheng, Y. M. Effects of super-absorbent polymers on a soil–wheat (Triticum aestivum L.) system in the field. Appl. Soil. Ecol. 73, 58–63 (2014).Article 

    Google Scholar 
    Li, R., Hou, X., Li, P. & Wang, X. Multifunctional superabsorbent polymer under residue incorporation increased maize productivity through improving sandy soil properties. Adv. Polym. Technol. 2022, 1–12 (2022).Article 
    ADS 

    Google Scholar 
    Singh, I., Verma, R. R., Srivastava, T. K. & Growth Yield, irrigation water use Efficiency, juice quality and economics of sugarcane in Pusa hydrogel application under different irrigation scheduling. Sugar Tech. 20, 29–35 (2018).Article 

    Google Scholar 
    Jat, A. L., Rathore, B. S., Desai, A. G. & Shah, S. K. Production potential, water productivity and economic feasibility of Indian mustard (Brassica juncea) under deficit and adequate irrigation scheduling with hydrogel. Indian J. Agricultural Sci. 88, 212–215 (2018).Article 

    Google Scholar 
    Saini, A. K., Patel, A. M., Chaudhary, P. P. & Saini, L. H. Impact assessment of Irrigation, fertility and hydrogel levels on growth Attributes, yield and economics of summer Pearl millet (Pennisetum glaucum L.) under North Gu-jarat conditions. J. Pharmacogn Phytochem. 7, 2914–2918 (2018).
    Google Scholar 
    Download referencesFundingThe work was supported by the projects of Technology Agency of the Czech Republic SS06020468 „Development of natural nutrient-releasing controlled-release hydroabsorbents for use in crop production”.Author informationAuthors and AffiliationsDepartment of Agrochemistry, Soil Science, Microbiology and Plant Nutrition, Faculty of AgriSciences, Mendel University in Brno, Zemědělská 1, Brno, 61300, Czech RepublicTomáš Kriška, Jiří Antošovský, Martin Brtnický, Jiří Kučerík, Jiří Holátko & Petr ŠkarpaInstitute of Materials Science, Faculty of Chemistry, Brno University of Technology, Purkyňova 118, Brno, 61200, Czech RepublicJosef JančářAuthorsTomáš KriškaView author publicationsSearch author on:PubMed Google ScholarJiří AntošovskýView author publicationsSearch author on:PubMed Google ScholarMartin BrtnickýView author publicationsSearch author on:PubMed Google ScholarJiří KučeríkView author publicationsSearch author on:PubMed Google ScholarJiří HolátkoView author publicationsSearch author on:PubMed Google ScholarJosef JančářView author publicationsSearch author on:PubMed Google ScholarPetr ŠkarpaView author publicationsSearch author on:PubMed Google ScholarContributionsTK was involved in conceptualization, investigation, data curation, software, writing—original draft. JA was involved in investigation, writing—review and editing. MB was involved in investigation, data curation, validation, and writing—review and editing. JK was involved in writing—review and editing. JH was involved in writing—review and editing. JJ was involved in methodology, and investigation. PS was involved in conceptualization, methodology, formal analysis, funding acquisition, supervision, software, writing—original draft. All authors read and approved the final manuscript.Corresponding authorCorrespondence to
    Petr Škarpa.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Ethics approval and consent to participate
    The cultivation of Papaver somniferum for experimental purposes was conducted in accordance with Czech national regulations (Act No. 167/1998 Coll. on addictive substances). The field trial at the Žabčice experimental station was duly reported to the competent authority according to the regulations.

    Consent for publication
    All authors give permission to the publisher to publish this research work.

    Additional informationPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary InformationBelow is the link to the electronic supplementary material.Supplementary Material 1Rights and permissions
    Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
    Reprints and permissionsAbout this articleCite this articleKriška, T., Antošovský, J., Brtnický, M. et al. Comparative effects of synthetic and natural hydrogels enriched with fertilizer on poppy yield and soil health in drought-prone conditions.
    Sci Rep 15, 44694 (2025). https://doi.org/10.1038/s41598-025-28213-0Download citationReceived: 07 July 2025Accepted: 10 November 2025Published: 29 December 2025Version of record: 29 December 2025DOI: https://doi.org/10.1038/s41598-025-28213-0Share 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
    KeywordsHydrogelCulinary poppyYieldSoil healthProfitability of fertilization More

  • in

    Odour learning concentration influences concentration range of conditioned response in newborn rabbits

    AbstractNewborn mammals must adapt to the chemically complex environment by detecting and prioritizing relevant stimuli. In rabbits, the mammary pheromone (MP) emitted by lactating females triggers a typical behaviour in newborns, helping them to locate the nipples, suck and survive. The MP also promotes very rapid learning of other odorants by associative conditioning. In this study, MP-induced learning was used to investigate the neonatal detection and recognition abilities of two odorants very different in volatility, ethyl isobutyrate and ethyl maltol, across concentrations ranging from 10− 5 to 10− 25 g/ml. The results show, firstly, that the odorants could be learned even at very low concentrations; and secondly, that a process of generalisation of the odorant quality was effective after learning over a wide range of concentrations. However, the degree of generalisation depended on the concentration at which the odorants had been learned, with quality and intensity becoming closely interdependent for very low concentrations of learning. Taken together, these data highlight the remarkable adaptability of the olfactory perceptual and cognitive systems of newborn rabbits, enabling them not only to rapidly learn new odorants, but also to attribute qualities to them that depend on the quality perceived at the learning concentration.

    Similar content being viewed by others

    Noradrenergic modulation of pheromone-induced odor learning and brain activation in newborn rabbits

    Article
    Open access
    28 November 2025

    Predicting odor from molecular structure: a multi-label classification approach

    Article
    Open access
    16 August 2022

    Data-science based analysis of perceptual spaces of odors in olfactory loss

    Article
    Open access
    19 May 2021

    IntroductionPerception occurs when animals create brain representations of the external world. It is critical in helping organisms make important choices among the many, varied and ever-changing sensory stimuli in the environment. This requires animals to discriminate between stimuli according to their nature but also their intensity, variations that can call for perceptual discrimination or, on the contrary, generalisation1,2. Thus, the power and adaptive capacity of sensory receptors, as well as the depth of neurophysiological processing of sensory inputs determine the way in which organisms interact with sensory stimuli. This shapes the perception of their own world (i.e., their Umwelt)3. Much remains to be discovered about the abilities of animals to adapt to changing environments as well as the cognitive and physiological limits underlying these adaptations, particularly early in life.Animals are regularly confronted with ecological situations that push their perceptual channels beyond their routine4. This can involve extreme sensitivity to odours in order to cope with critical situations such as, in adult organisms, finding food, avoiding predators, choosing a mate, caring for offspring5,6,7,8,9; for example, adult male insects can detect odour molecules at remarkably low concentrations to track the pheromone plumes emitted by females10,11. Importantly, olfaction is also a major sensory feature of young organisms, contributing to the adaptation of individuals to the environment early in ontogeny12. This raises the question of the extent to which organisms have effective and sufficiently efficient olfactory detection abilities at birth, and how these performances can be modulated by learning, when the neonatal olfactory system and brain are still partially immature. Here, we tested these research questions in an animal model that has become very relevant in recent decades for assessing both predisposed and learned perinatal odour perception: the newborn rabbit.In mammals, odour perception begins in utero and is vital from birth when newborns must immediately adapt to the external (aerial) environment12,13,14. Neonatal responsiveness to odours may result from predisposition to respond to spontaneously significant biological stimuli, and/or from learning about stimuli that were initially devoid of meaning15,16,17. The rabbit model illustrates these two mechanisms. Indeed, newborn rabbits are remarkably quick to locate their mother’s nipples during the single and very brief (< 5 min) daily exposure to the mother in the nest, even though they are initially blind and deaf. This is due to their perception of a monomolecular volatile component naturally emitted by lactating female rabbits in their milk, known as the mammary pheromone (MP, 2-methylbut-2-enal)18,19,20,21,22. The MP triggers the typical orocephalic nipple-search behaviour in newborns, without any learning beforehand19,23. The MP also acts as a potent reinforcer, i.e. an unconditioned stimulus (US), which can rapidly confer biological significance to a new odorant/mixture of odorants (conditioned stimulus, CS) that is initially inactive (i.e., not spontaneously active in behaviour). Twenty-four hours after a single exposure to CS + US for a very brief presentation period (5 min) the CS alone is sufficient to trigger the typical sucking-related behaviour in newborn rabbits24,25,26,27,28.It has recently been shown that a single episode of MP-induced odour learning is sufficient to increase the strength of electrophysiological responses to the CS in the olfactory mucosa of rabbit pups, and their perceptual sensitivity to the CS29. This was the first evidence of the process of “induction” in neonates, i.e., when odour conditioning induces plasticity not only in the brain but also in the periphery of the olfactory system30,31,32,33,34,35. Thus, whereas newborn rabbits could not perceive a new odorant at 10− 17 g/ml because this concentration was below their spontaneous detection threshold for that odorant, they could respond to it at 10− 17 g/ml 24 h after learning the odorant by pairing with the MP; this was observed when both stimuli were used at 10−5 g/ml during conditioning. Therefore, MP-induced odour learning can not only confer biological value to a new odorant, but also improve the newborn’s sensitivity to that odorant by expanding their detection threshold to lower concentrations29.In the present study, we aimed to confirm and extend these recent observations while examining in depth how the processes of generalisation and discrimination interact in neonatal odour detection. The experiments were conducted in 2- to 4-day old rabbit pups, using two odorants known to be learnable by newborn rabbits26,27,29. These odorants were deliberately chosen for their contrasting volatility properties – odorant A (ethyl isobutyrate) is highly volatile, whereas odorant B (ethyl maltol) is low in volatility – in order to reflect the natural environment, which contains odours that vary widely in volatility. First, we conditioned rabbit pups to odorant A or B, with MP concentration maintained at a constant level (10− 5 g/ml) while the CS concentration varied over a wide range, i.e., from 10− 3 to 10− 24 g/ml. The hypothesis was that MP could promote the learning of CS even at very low concentrations during the conditioning, with a certain limit (to be determined). Second, the responsiveness of the pups after MP-induced learning was systematically assessed not only by testing at the CS concentration to which the pups had been conditioned, but also by presenting them with a wide range of CS concentrations, including levels higher and lower than the initial CS concentration. Our hypothesis was that the concentration range within which newborn rabbits respond to CS after conditioning would vary according to the concentration of CS during conditioning, i.e., that memory may depend on the quality perceived and acquired during conditioning.ResultsPup responsiveness to conditioned concentrationsTo assess the concentration range over which rabbit pups can be conditioned to a new odorant by pairing it with MP 10− 5 g/ml, 176 and 153 pups were conditioned on day 1, 2 or 3 to odorant A from 10− 5 to 10− 23 g/ml (9 groups, one per concentration) and odorant B (8 groups, one per concentration) from 10− 5 to 10− 22 g/ml, respectively, before being tested to the conditioned stimulus 24 h after the conditioning episode. Analysis of the results then consisted of examining the responsiveness displayed to the concentration at which the CS had been learned.For odorant A, the pups highly and similarly responded to the CS when its learning concentration was between 10− 5 and 10− 22 g/ml (77–100%; χ² = 9.1; df = 7, p = 0.24; all statistical results are detailed in supplementary files, only the main results are presented in the text). At these concentration levels, their responsiveness was as high as to the MP itself (95–100% response to MP; 2 × 2 comparisons between learning concentrations of A vs. MP: χ² < 0.5, p > 0.24). However, pups did not respond to odorant A after conditioning at 10− 23 g/ml (2 × 2 comparisons with all other concentrations: χ² > 20.49, p < 0.0001), whereas they responded strongly to MP (100%; not illustrated here) (Fig. 1A).Fig. 1Proportions of 2- to 4-day-old rabbit pups responding in an oral activation test to (a) odorant A (ethyl isobutyrate, n = 176 pups) and (b) odorant B (ethyl maltol, n = 153 pups) at the concentration at which each stimulus has been conditioned by pairing with the mammary pheromone 24 h before, i.e. 10− 5 to 10− 23 g/ml for A, and 10− 5 to 10− 22 g/ml for B.Full size imageFor odorant B, the pups highly and similarly responded to the CS when its concentration of learning was between 10− 5 and 10− 21 g/ml (65–95%; χ² = 6.6, ddl = 6, p = 0.35). Again, this responsiveness was as high as to the MP itself (89–100%; 2 × 2 comparisons between learning concentrations of B vs. MP: χ² < 3.2, p > 0.05). However, their responsiveness to odorant B was null at 10− 22 g/ml (2 × 2 comparisons with all other concentrations: χ² > 15.51, p < 0.0001) whereas the pups responded to MP (100%; not shown here) (Fig. 1B).In addition, we tested whether newborn rabbits can learn the odorants A or B at concentration levels higher than 10− 5 g/ml, i.e., at 10− 3 and 10− 2 g/ml. A prerequisite was to determine whether the MP could be detected in such mixtures. Therefore, 10 naive 2-day-old pups were tested for their oral responsiveness to two mixtures containing MP at 10− 5 g/ml and odorant A at 10− 2 or 10− 3 g/ml. Nine other naive 2-day-old pups were tested for their oral responsiveness to 2 mixtures containing MP at 10− 5 g/ml and odorant B at 10− 2 or 10− 3 g/ml. For odorant A, no pups responded to the mixtures, suggesting that the MP was masked by odorant A, making learning impossible at these concentrations. Indeed, when 5 and 9 pups were exposed to MP + A 10− 2 or MP + A 10− 3 g/ml, respectively, none of them responded to odorant A the following day (while all responded to the MP). In contrast, 90% and 100% of pups responded to the MP mixed with B at 10− 2 and 10− 3 g/ml, respectively, demonstrating that MP was perceived in these mixtures and, thus, that B could theoretically be learned. However, when 5 and 10 additional pups were exposed to MP + B 10− 2 or MP + B 10− 3 g/ml, respectively, only one pup conditioned with B at 10− 3 responded to B on day 3 (while all responded to MP) demonstrating that odorant B cannot be learned efficiently at these concentrations.In summary, this first set of results showed that, under our conditions and with a single concentration of MP, newborn rabbits could learn odorants A or B over a wide range of concentrations (of 1017 and 1016 log-units, respectively). This range included very low concentrations (< 10− 20 g/ml for A and B; with certain limits: no learning occurred below 10− 21 g/ml for B and below 10− 22 g/ml for A) but excluded very high concentrations (≥ 10− 3 g/ml for both odorants).Pup responsiveness to different CS concentrations as a function of conditioning concentrationTo determine how MP-induced conditioning gave rise to a qualitative generalization and/or an increase in sensitivity depending on the learning concentration, the pups conditioned above to odorants A or to B were also tested on concentrations other than the learning ones. To this end, each group of pups was divided into two subgroups tested with 5 or 6 concentrations of CS (from 10− 5 to 10− 25 g/ml and from 10− 5 to 10− 22 g/ml for A and B, respectively) systematically including concentrations expected to be below the detection/recognition threshold, then with the learning concentration, and finally with the MP as a control. Analysis of the results consisted of examining the responsiveness of the pups within and between the subgroups.For odorants A and B, whatever the concentration of learning, the distribution of responsiveness over the ranges of concentrations differed according to the tested concentrations (global comparisons for A and B: p < 0.043 for A and p < 0.002 for B – with rare exceptions for CS B− 17 and B− 19, p = 0.1 -; detailed statistics are provided in supplemental materials). Below, we detail the pups’ responsiveness when it was statistically significant (all comparisons, including these, are presented in the supplementary materials).For odorant A (Fig. 2, left), conditioning at 10− 5 g/ml was followed by high and constant responsiveness to the stimulus (> 60%) from 10− 5 to 10− 18 g/ml. Then, a drastic drop was observed at 10− 19 g/ml and responsiveness remained very weak (≤ 20%) at the lowest tested concentrations (10− 20, −21, −22 g/ml). After conditioning at 10− 10 g/ml, only 50% of the pups responded at the highest concentration (10− 5 g/ml) while 67–90% responded over a range extending from 10− 10 to 10− 21 g/ml; this was one of the widest ranges of responsiveness observed in our results despite a transient slight decrease at 10− 20 g/ml. The responsiveness then severely dropped at 10− 22 g/ml. After conditioning at 10− 15 and 10− 17 g/ml, pups completely ceased to respond to the highest concentrations, i.e., from 10− 5 to 10− 10 and 10− 5 to 10− 15 g/ml, respectively. The range of active concentrations became limited to 10− 15 to 10− 18 g/ml (CS A 10− 15 g/ml: ≥ 80%) and 10− 16 to 10− 21 g/ml (CS A 10− 17 g/ml: ≥ 67%), with, for the latter, a relative drop at 10− 20 g/ml (40%), whereas 10− 21 induced 78% responsiveness. After conditioning at 10− 19 g/ml, the highest responsiveness was observed between 10− 18 and 10− 21 g/ml (≥ 75%), while responsiveness was intermediate at 10− 17 and 10− 22 g/ml (42–50%), very weak or null for 10− 16 and above, and 10− 23 g/ml and below. For 10− 20 g/ml of learning concentration, high responsiveness was observed from 10− 19 to 10− 24 (60–91%), but not significantly above or below these concentrations (< 20%). After conditioning at 10− 21 and 10− 22 g/ml, the range of concentrations triggering the highest responsiveness became narrower and centred on the learning concentration, i.e., from 10− 20 to 10− 22 g/ml (90–100%) and from 10− 21 to 10− 23 g/ml (55–86%) respectively. Finally, after conditioning at 10− 23 g/ml (not shown in Fig. 2), no pups responded to A 10− 23 g/ml (Fig. 1A) or to any of the other concentrations tested (from 10− 18 to 10− 25 g/ml), attesting that the learning concentration threshold for odorant A is ≤ 10− 22 and > 10− 23 g/ml.Fig. 2Proportions of 2- to 4-day-old rabbit pups responding in an oral activation test to the conditioned odorant A (ethyl isobutyrate, black bars, n = 162 pups) or odorant B (ethyl maltol, grey bars, n = 138 pups) 24 h after MP-induced conditioning to the conditioned stimulus (CS) at concentrations varying from 10− 5 to 10− 22 g/ml for A, and from 10− 5 to 10− 21 g/ml for B. For each learning concentration of CS, i.e., each graph, two distinct groups of pups were tested at various levels of concentrations during oral testing. In each graph, responsiveness to the CS concentration used at conditioning is marked by a white bar and, significant response rates between the highest and lowest levels of responsiveness are marked with a star (detailed statistical comparisons are provided in Supplementary Materials).Full size imageFor odorant B, we observed the same global trend as for odorant A in pup responsiveness, but with a different dynamic. After conditioning at 10− 5 g/ml, responsiveness was high (≥ 80%) over a wide range of concentrations, i.e., from 10− 5 to 10− 19 g/ml, then severely decreased at 10− 20 g/ml (22%) and completely ceased below this concentration. After conditioning at 10− 10 and 10− 15 g/ml, pups responded highly over a narrower range, anchored along the highest concentrations, i.e., from 10− 5 to 10− 15 g/ml (84–89%) and 10− 5 to 10− 16 g/ml (70–89%), respectively, while responsiveness was weak at lower concentrations (≤ 30%). After conditioning at 10− 17 g/ml, the range of pup responsiveness became again considerably broader: pups responded over a range of concentrations extending to the lower concentrations, i.e., from 10− 15 to 10− 21 g/ml (40–70% of responsiveness). Thus, the reactive concentration range was as broad as for B 10− 5 g/ml but with lower response rates. From conditioning at 10− 19 g/ml, pup responsiveness became more and more narrow, restricted to the lowest concentrations, and almost exclusively displayed only to the learning concentration or to adjacent, very similar concentration levels. Thus, after conditioning at 10− 19 and 10− 20 g/ml, pups responded to B from 10− 18 to 10− 21 (40–70%) and at 10− 17, −19, −21 g/ml (40–65%), respectively. After conditioning at 10− 21, they responded only to B 10− 20 and 10− 21 g/ml (64–76%). Finally, after conditioning at 10− 22 g/ml, no pups responded to B 10− 22 g/ml (Fig. 1B) or to any of the other concentrations tested (from 10− 18 to 10− 23 g/ml), demonstrating that they could no longer learn odorant B from 10− 22 g/ml, and attesting that the learning concentration threshold for B is ≤ 10− 21 and > 10− 22 g/ml.In summary, for both odorants A and B, the pups were able to learn the CS over a wide range of concentrations – with certain limits -, and the learning concentration always induced one of the highest levels of responsiveness. In addition, two main results emerged that were similar for both odorants. First, for the highest concentrations of learning, i.e., 10− 5 g/ml for B and 10− 5 and 10− 10 g/ml for A, behavioural responsiveness showed the most extended generalization, with very low detection/recognition thresholds. Second, for the lowest learning concentrations, the active concentration range shifted to lower concentrations and became narrower around the learning concentration. Thus, the results indicate that the newborn rabbits learned a concentration-dependent odour quality for each odorant, meaning that the learning concentration played a critical role in the newborns’ recall and post-conditioning perceptual abilities.Responsiveness to extremely low levels of concentration and order effectThe responsiveness of newborn rabbits to the lowest concentrations tested (below 10− 20 g/ml) raised questions about the functioning of olfaction, i.e., how molecules can sufficiently enter into the nares and interact with receptors to generate perception when there are so few of them (see General discussion). Therefore, we wanted to ensure that this responsiveness was not the result of a methodological factor. Indeed, the order in which different stimuli are tested can theoretically affect how organisms respond to them. Here, testing the same rabbits at several different concentrations in succession could generate responses linked to an order effect. For example, exposure to a high concentration first could result in a response to a very low concentration later on. We took several precautions to avoid this, such as presenting the concentrations in a different order for each newborn from a same litter, leaving a sufficient time interval between each presentation to the same pup, and conducting blind testing (see the Methods section for details). However, to ensure that there was definitely no order effect, we have tested 39 other newborn rabbits (from 6 litters) as follows: (1) 14 pups were conditioned exactly as before (i.e., by single pairing with MP at 10− 5 g/ml) to odorant A at 10− 21 g/ml; (2) 15 pups were conditioned to odorant A at 10− 22 g/ml; (3) 10 pups were conditioned to odorant A at 10− 23 g/ml (negative control group). Twenty-four hours later, each of the three groups was tested at the conditioning concentration only, followed by testing at MP 10− 5 g/ml. If there were an order effect, the levels of responsiveness to the odorant A (i.e., the proportions of pup responding) would differ from those observed in Fig. 2. Conversely, if the pups’ responsiveness remained the same, this would demonstrate that successive presentations do not affect them.As a result, 93%, 87% and 0% of the pups responded to A 10− 21, 10− 22 and 10− 23 g/ml, respectively, 24 h after conditioning. These results were not significantly different to those observed after successive testing of different concentrations of A (100%, 85%, 0%, respectively in Fig. 2; χ² < 0.5, p > 0.05). At the same time, 100% of the pups responded to MP.Thus, the responsiveness displayed by the rabbit pups in our experiments at extremely low concentrations did not appear to be due to an order effect.DiscussionThe aim of the present study was to investigate whether, when newborn rabbits learn a new odorant, the concentration at which the odorant is learned influenced the qualitative generalization of the learned stimulus, i.e., the range of concentrations to which the pups subsequently responded; and, if so, whether this differs as a function of the concentration level (high or low) at the time of learning. We evaluated this by harnessing the powerful ability of newborn rabbits to learn novel odorants through direct pairing with the naturally relevant mammary pheromone (associated with milk intake and survival in this species). The results largely extend our initial observations29, by applying the same procedures but with two odorants and over a much wider range of concentrations.Under our experimental conditions, newborn rabbits exhibited remarkable sensory and cognitive abilities (which we discuss in detail below). Under these rigorously controlled conditions (see the Methods section), the conditioned odours elicited neonatal responsiveness at certain concentrations but clearly not at others, a result consistently obtained with two odorants of opposite volatility. This indicates, firstly, that our procedures reveal not only the remarkable abilities of rabbit pups, but also the existence of some limitations in their ability to detect and integrate novel stimuli; and secondly that the responsiveness of the pups was due solely to the presented stimuli and not to contextual artefacts that might interfere with them.At the highest concentrations, exceeding 10− 5 g/ml, newborn rabbits were unable to learn the odorants A (highly volatile) and B (weakly volatile) in our experimental conditions (i.e. with MP used as the unconditioned stimulus at 10− 5g/ml in mixture with A or B). For A at 10− 2 and 10− 3 g/ml, behavioural non-reactivity of naive (i.e., unconditioned) newborns to the A + MP mixture revealed that odorant A masked the perception of the MP (A was then about 6000 and 600 times more concentrated than the MP at 10− 5 g/ml). This masking prevented association and learning36,37. In contrast, for odorant B at 10− 2 or 10− 3 g/ml, naive newborn rabbits responded to B + MP, showing that odorant B did not mask the perception of the MP (given its low volatility, B was 2000 and 200 times less concentrated than MP). Nevertheless, learning of B failed to take place. This lack of learning may be due to the amount of alcohol required to make B soluble (B was poorly soluble at these concentrations and required 1 and 0.1% of alcohol at 10− 2 and 10− 3 g/ml). The alcohol may irritate the pups and/or produce an unpleasant sensation, thus interfering with the learning of odorant B.From 10− 5 g/ml and below, we discuss the responsiveness of newborn rabbits in four points.First, rabbit pups learned the odorants along a particularly wide range of concentrations. Indeed, rabbit pups learned odorant A down to 10–22 g/ml, and odorant B down to 10–21 g/ml. These learnable concentrations are de facto the lowest concentrations that could have been perceived on the day of co-exposure with the MP. They thus correspond to the spontaneous detection thresholds of odorants A and B in our conditions. Knowing that these results hold both for a naturally high (A) and low volatile odorant (B), one may assume that they can be generalised to most odorants. This highlights the capability of newborn rabbits to detect and learn novel stimuli in their environment even at impressively low concentrations, at least in association with the major biological signal spontaneously provided by the MP in this species. It is also important to note that the MP imparts biological significance on odorants A and B at concentrations at which it is not itself spontaneously active as a behavioural releaser (active range of MP as releaser in newborn rabbits: 2.5 × 10–5 to 2.5 × 10–9 g/ml [38]).Second, the pups showed the widest continuous range of responses for the highest learnable concentrations of CS, i.e., 10− 5/10− 10 g/ml for A and 10− 5 g/ml for B. Indeed, for these learning concentrations, the pups were able to respond after conditioning to odorant A down to 10− 21 g/ml and after conditioning to odorant B down to 10− 19 g/ml, i.e., up to 17 and 15 log-units lower than the concentration at learning. These results highlight that even early in life, animals can generalise a learned stimulus quality when it is subject to large quantitative variations, ensuring the perception of an odour quality continuum; however, this process has certain limitations, probably physiological and cognitive, as can be seen from our results. In animals – including humans – generalisation is a fundamental feature of odour-guided adaptive behaviours38,39,40,41 and most researchers agree that it constitutes a priority in the general functioning of the olfactory system42,43,44,45,46. To meet this priority, our results suggest that, spontaneously, the rabbit olfactory system commonly operates within a concentration range limited to lower concentrations of 10− 10 for A and 10− 15 g/ml for B. Above this limit, i.e. over a range of concentrations encompassing intermediate and high concentrations, odour identity might be spontaneously invariant46. However, the MP-induced learning could push back this limit and allow much wider generalisation. This would be particularly useful in fasted newborns, as here, i.e., in pups which are in a state of maximum wakefulness47,48.Third, at learning concentrations lower than 10− 10 and 10− 15 g/ml for odorants A and B, the pups did not respond to the highest concentrations during behavioural testing, revealing a disruption in the perceived quality of the stimulus. From this disruption, the distribution of behavioural responsiveness indicates that new quality/qualities would emerge, inextricably dependent on intensity. This could create a generalisation gradient around the learned concentration, at which the qualitative attributes of the stimulus are optimally recognised. Below and above the learned concentration, the proportion of responding pups decreased, most likely in proportion to the perceived ‘dissimilarity’ in stimulus quality49,50. Thus, if a pup was conditioned to a high concentration, it generalized to all concentrations down to a very low one. A question might then be: why doesn’t it generalize to the same low concentrations when conditioned to a slightly lower concentration? At high concentrations, learning process probably implies that almost all the receptor neurons sensitive to the odorant were recruited and thus stimulated51. In contrast, at low concentrations, learning probably involved stimulation of a much smaller subset of receptor neurons (only the most sensitive)52. This hypothesis could explain the observed asymmetry in generalization. Indeed, at high concentrations of CS, the peripheral activity elicited by lower concentrations during post-conditioning tests would correspond to only a fraction of the activity expressed and encoded during learning; however, from this fraction, the phenomenon of ‘pattern completion recognition’1,53 could come into play, allowing the animal to recognize the stimulus as similar to CS. Conversely, at low concentrations of CS, the peripheral activity generated by higher concentrations has never been induced before, and never memorized, so the phenomenon of pattern extension from the encoded fraction cannot occur. In any case, the results showed that, at low concentrations, the pups then learn a quality more dependent on the concentration of the odorant – change the concentration and the quality changes such that during retention test there is no congruence, and no response. So there would be a quality/intensity tuning related to the concentration. The existence of such a tuning may have important implications for the design of learning protocols aimed at detecting odorants in natural olfactory scenes; the protocols should be adapted to the expected target concentrations54. This concept may open new insights on the coding of olfactory quality in general and stimulate further work to explore the neurophysiological basis of this coding.Fourth, the lower the CS concentration at learning, the lower the detection/recognition threshold. Thus, when odorant A was learned at concentrations from 10− 20 to 10− 22 g/ml, and odorant B from 10− 19 to 10− 21 g/ml, the after-learning detection threshold was 10− 24 g/ml and 10− 21 g/ml, respectively. Importantly, these concentrations were not learnable by rabbit pups (i.e., cannot be used as CS) but became perceived after MP-induced conditioning. Thus, such conditioning not only increased the operating range of the olfactory system and thus qualitative generalisation, but also can slightly increase the olfactory detection performance of rabbit pups. The peripheral plasticity may explain, at least in part, this gain in sensitivity29. How do the detection thresholds we have identified in newborn rabbits compare with the lowest thresholds reported in the mammalian literature? We propose here a comparison, with the caveat that the experimental conditions are not identical (e.g., not at the same developmental stage of organisms, not with the same odours, the same experimental/learning procedures, etc.). For instance, after conditioning, adult mice can detect bourgeonal at 10− 4 ppt55 and isoamyl acetate at or below 1 ppt56. In adult dogs, the reported post-conditioning behavioural threshold for amyl acetate was 1–2 ppt57,58. In newborn rabbits, using odorant A, whose volatility properties are closest to those of isoamyl acetate and bourgeonal, the post-learning detection threshold was 10− 24 g/ml, equivalent to approximately 10− 8 ppt (g/g) (Table 1). Thus, newborn rabbits would have a sensitivity 104 to 108 times greater than those of adult mice and dogs. This may be because newborn mammals need to quickly learn about their environment, and that their developing brains are particularly well suited to forming associations easily. As mentioned above, this difference may also be partly due to our experimental conditions. Indeed, we worked with rabbit pups: (a) in a learning paradigm using an unconditioned stimulus, the MP, which is of very high spontaneous biological value, essential for newborn rabbits’ survival; (b) the same stimulus (MP) was used to give a conditioned stimulus the power to release the behavioural response initially elicited by the MP alone; (c) we compared the responsiveness to the CS after the pups had learned it at different concentrations. In comparison, studies in mice and dogs did not use a pheromone as an unconditioned stimulus and tested the consequences of learning a CS at only one concentration during conditioning.Table 1 Concentrations of stimuli in aqueous phase. Odorant A: Ethyl isobutyrate (CAS# 97–62–1), MW (molecular weight) = 116.16 g. Odorant B: Ethyl maltol (CAS# 4940–11–8), MW = 140.14 g. Concentration in ppt (part per trillion g/g) = C mol/l x MWg x 1012.Full size tableTogether, the lowest learnable and post-learning detectable concentrations observed in our study demonstrate that rabbit pups have a very high odour sensitivity. Although the results were unexpected at low concentrations, their remarkable consistency and reproducibility, combined with our methodological controls and validations, give them scientific validity. However, we are fully aware of the extremely low theoretical concentrations of A and B sometimes used, particularly below 10− 20 g/ml. Given the current state of knowledge regarding olfaction, we are not yet able to explain how a very small number of molecules can generate the detection and perception of a signal in rabbit pups at such low concentrations. To facilitate comparison with some values of olfactory sensitivity reported in the literature, Table 1 shows the liquid-phase concentrations of A and B expressed in mol/L and ppt (g/g). It should be noted that these values are estimated based on serial dilution calculations, and that the actual concentrations of the stimuli in the gaseous phase are unknown. To date, to the best of our knowledge, no instrument is sensitive enough to measure such concentrations directly.Although the literature does not report such concentration values by name, olfactory sensitivities of this order may be involved in the behaviour of other aerial-breathing animals in nature. Indeed, the lowest concentrations used here may correspond to the conditions that animals encounter in open spaces when detecting or following either scent trails, i.e., stimuli that have become biologically relevant through experience (learning), or spontaneously relevant stimuli such as pheromones59. For example, after learning artificial odours under experimental conditions, dogs60,61 or butterflies in the wild62 can follow olfactory trails/pheromonal plumes for several kilometres. In nature, thousands of m3 of air dilute olfactory trails, even if one supposes that the heaviest molecules will be deposited on the ground or distributed according to an organized ‘moving plume’. As an illustration, Bombyx Mori males can detect as few as 2.8 × 10− 22 mol/l in the gradient of molecules of bombykol released by females, which constitute the pheromone plume; only one pheromone molecule is then needed to elicit a nerve impulse in olfactory sensory neurons11. Remarkably, the capabilities observed in newborn rabbits during our experiments are at the same order as those observed in Bombyx mori, with rabbit pups’ detection thresholds of 8.61 × 10− 24 mol/l for A and 7.14 × 10− 21 mol/l for B (Table 1). Although this comparison involves an insect and a mammal, as well as direct responsiveness to a pheromone versus to an odorant learned through pheromone-induced conditioning, the shared factor of pheromone signaling bridges this phylogenetic gap: it directs mate-searching in silk moths and food-searching in newborn rabbits – two behaviors critical for survival. In both cases, pheromones form the basis of intense behavioral motivations that probably prompt the olfactory system and other sensory systems to function under extreme conditions.In conclusion, this study shows that mammary pheromone-induced learning in newborn rabbits produces a remarkable sensitivity to the learned odorant, a sensitivity that varies according to the concentration at which the odorant was learned. Notably, these effects are observed whether the conditioned stimulus is an odorant of high or low volatility. From an ecological perspective, the remarkable ability of newborn rabbits to learn and respond to new information, depending on the chemical context in which learning occurs, is adaptive. In the wild, the nature and intensity of the odorants carried and emitted by rabbit mothers can vary between females, probably depending on factors such as age, physiological and social status, and diet. It is therefore crucial for newborns to be able to adapt to this variability in order to suck efficiently when the mother visits the nest (for only five minutes per day), and to learn about their mother’s environment and that of their social group, which they will encounter when they are old enough to leave the nest. In terms of generalisation, the results show that if newborns are exposed to relatively high stimulus concentrations during learning, they subsequently demonstrate broad qualitative generalisation with respect to the conditioned stimulus; conversely, if they are exposed to low concentrations during learning, they subsequently respond only to concentrations closely tuned to the learned concentration. In the first case, this may be a way for newborns to learn information that is widespread in their environment and that it will be useful for them to recognise later on, even when it undergoes variations in chemical intensity (e.g., odors from familiar individuals in their own or other social groups, or from familiar food, which may vary according to the seasons). In the second case, it may facilitate the identification/recognition of highly specific information that must be clearly distinguished from other information (e.g., odors allowing the recognition of the mother from other lactating females, or the perception of non-social signals that are crucial for survival, such as danger cues).Taken together, the results confirm that newborn rabbits are a particularly valuable model for exploring the functioning of olfaction, including in conditions of concentration that have rarely or never been studied before in aerial and aquatic animals10,11,63. Studying their olfactory capacities will allow researchers to further uncover the fascinating perceptual and biological abilities displayed by animals, particularly during the early stages of life.MethodsEthicsThis work is reported in accordance with ARRIVE guidelines and strictly followed the local, institutional and French national rules regarding the care and experimental uses of the animals. Thus, all experiments were conducted in accordance with ethical guidelines enforced by French Law (French Ministries of Agriculture, and of Research & Technology) and approved by the Ethical Committees for Animal Experimentation of the University of Lyon 1 (CEEA-42 and − 55) and the French Ministry of Higher Education and Research (no. APAFIS #27874–2020110416356847 v2).AnimalsNew-Zealand rabbits (Charles River strain, France) originated from the breeding colony of the Centre de Recherche en Neurosciences de Lyon. 12 adult males and 40 adult females Oryctolagus cuniculus (Charles River strain, L’Arbresle, France) were kept in individual cages under a constant 12:12 light-dark cycle (light on at 7:00 a.m.), with ambient air temperature maintained at 21–22 ◦C. Water and pelleted food (Lapin Elevage 110, Safe, France) were provided ad libitum. The study, which lasted over two years (19 experimental sessions were conducted), used two consecutive parent herds, one before and one after the peak of COVID pandemic in France. Two days before the expected day of parturition (day of delivery was considered postnatal day 0; d0), a nest box (0.39 × 0.25 × 0.32 m) was fixed to the cages of pregnant females. To even out pup-female interaction, females’ access to the nest was allowed for 15 min per day at 11:30 a.m. (this procedure allowed mimicking the short daily nursing displayed by rabbit females64. We used 387 pups of d1-4 from 67 litters.OdorantsThe odorants consisted of 2-methylbut-2-enal (the Mammary Pheromone, MP, CAS# 497-03-0)19,20,21,22,23,24,25,26,27,28,29, ethyl isobutyrate (odorant A; CAS# 97-62-1) and ethyl maltol (odorant B; CAS# 4940-11-8) (purity of the odorants ≥ 99%)26,29,66. The odorants A and B had very high and low volatility properties, respectively, characterized by a factor of volatility approximately 167 000 higher for A compared to B (vapour pressure saturation: 3.23 × 103 Pa for A versus 1.93 × 10− 2 Pa for B). To give an idea in the gas phase, the ratio of A and B can be estimated using Henry’s law constants (as is standard practice in olfaction): A is approximately 1.3 × 10⁶ times more concentrated than B, at all dilution levels. Thus, the odorants A and B were representative of the opposite poles of the various volatilities of environmental chemical components, enabling us to generalise our observations.The MP allowed us to induce learning of odorants A or B through associative conditioning (see below, the odour conditioning section). The MP served as the unconditioned stimulus and was used at 10− 5 g/ml, a concentration known to be highly efficient in promoting conditioning24,25,29, while the A and B odorants served as the conditioned stimuli (CS). For A and B, the main set of concentrations mixed with the MP was 10− 5, 10− 10, 10− 15, 10− 17, 10− 19, 10− 20, 10− 21, 10− 22 g/ml, and 10− 23 g/ml for A only. In addition, in separate tests and for both odorants, two higher concentrations were also mixed with MP, 10− 2 and 10− 3 g/ml. During behavioural testing, MP was also used at 10− 5 g/ml and single odorants A and B were used at concentrations varying from 10− 5 to 10− 25 g/ml (odorant A), and from 10− 5 to 10− 22 g/ml (odorant B) were also used (see below).All the odorants were purchased from Sigma-Aldrich (Saint-Quantin Fallavier, France). The final solutions were obtained from stock solutions at 10− 2 g/ml prepared in 0.1% of ethanol (anhydrous, Labelians, Nemours, France) and distilled water; these stock solutions were renewed several times a year (in new vials). Then, successive dilutions were prepared in distilled water only, following dilution steps of 100, excepted for 10− 3 and 10− 15 g/ml, which were prepared following dilution steps of 10. For each experimental session, we prepared new series of dilutions in new vials. All preparations were made in a dedicated odorant storage and preparation room in our laboratory, under a hood.Odour conditioningConditioning sessions were run on days 1, 2 or 3 in an experimental room adjacent to the breeding room. These two rooms were isolated, each closed by a door, with no windows on the outside so no possible contamination from the outside. They were completely separate from the room where odour solutions were prepared (not on the same floor of our building). A ventilation system ensured constant and optimal air renewal in the two rooms. The pups were transferred from the breeding to the experimental room in groups of 5 (from the same litter) into a box lined with nest materials and maintained at room temperature (experimenters wore lab coats, hand gloves and no perfume on the days of the experiments). The MP-induced conditioning consisted of a single, brief and simultaneous exposure both to the MP and to the conditioned stimulus: 8 ml of the MP-A or MP-B mixtures were pipetted onto a cotton pad (19 × 14 cm, 100% cotton), then held 2 cm above the pups for 5 min. This exposure is known to induce very rapid learning of the stimulus paired with the MP24,25,26,27,28,29. The conditioning session took place 1 h before the daily nursing (10:30 a.m.), to standardise the pups’ motivational states and limit the impact of satiation on responses47. Two minutes after the end of the conditioning, the pups were individually marked with weakly odorous ink and returned to their nest (it was indeed important to be able to identify them individually until the next day – the day of behavioural test – because of the test procedure; see the section immediately below). The box containing the pups was rinsed with alcohol and distilled water after each conditioning session. Around 20 pups from 4 different litters were conditioned to the same concentration of the conditioned stimulus in order to control for individual differences; each of these groups were then split into two subgroups of 9 to 11 newborns (from 2 litters) for behavioural testing (see below).Behavioural assayBehavioural testing took place on days 2, 3 or 4, i.e., 24 h after phase 1 (conditioning), in the same experimental room as for conditioning (by experimenters again wearing lab coats, hand gloves and no perfume). It was also run 1 h before daily nursing to limit the impact of satiation on pups’ motivation and responsiveness47. The assay consisted of an oral activation test during which a pup was immobilized in one gloved hand of the experimenter, its head being left to move freely. Each CS (odorant A or odorant B) at different concentrations were presented for 10 s, as was the MP as a control at a constant concentration (10− 5 g/ml), using a glass rod held 0.5 cm in front of the nares, immediately after immersion in the diluted solution18,19,20,21,22,23,24,25,26,27,28,29. Each stimulus concentration had its own glass rod, which was rinsed/dried between animals and cleaned daily with alcohol.A test was positive when the CS elicited the stereotyped on/off response normally elicited by the MP, i.e., head-searching movements (vigorous, low amplitude horizontal and vertical scanning movements displayed after stretching towards the rod) usually followed by grasping movements (oral seizing of the rod extremity). Non-responding pups displayed no response except active sniffing. Pups were tested 5 by 5, so that the two subgroups of n = 10 allowed us to cover the range of concentrations evaluated for odorant A, as for odorant B. Each pup participated in only one experiment and was tested at a maximum of 5–6 concentrations of the conditioned stimulus (and to MP at the end) to prevent fatigue or habituation. Successive concentrations were tested blind to the experimenters. In practice, two experimenters were present at each test session and the one who actually performed the test with the rabbits (experimenter 1) did not know the concentrations he was testing. These concentrations were provided to him by experimenter 2 who held out the relevant vial without experimenter 1 being able to see the concentration indicated on it. The test consisted of the presentation of a first stimulus to a pup, its reintroduction into the box, then the stimulation of another pup with a second stimulus, its reintroduction into the box, and so on with all pups and all stimuli of a given series by respecting an inter-trial interval of 60 s. The order of presentation of the concentrations was strategically interspersed from one newborn to the next to prevent habituation/sensitization/extinction, except for the penultimate and the last stimulation, which were always the same, i.e. the CS at the learning concentration then the MP used as a control.When a pup responded to a stimulus, it usually made direct nasal contact with the end of the rod, often even grasping it in mouth, so its nose was gently dried with absorbent paper before the pup was returned to the box and exposed to further stimulation. The pups were immediately reintroduced to the nest after the end of their testing session, before a new group of 5 pups was tested in turn.Statistical analysisPup responding frequencies were compared using the χ² Pearson test when the groups or subgroups were independent (i.e., distinct groups/subgroups tested for their response to the same stimulus) or Cochran’s Q test when the groups/subgroups were dependent (i.e., pups from the same group/subgroup tested for their response to several stimuli). When the Cochran’s Q test was statistically significant, proportions of responding pups were compared 2 × 2 by the χ² McNemar test. Statistical analyses were performed using XLSTAT software (Microsoft, Redmond, USA). Degrees of freedom are indicated when > 1. Effects were considered significant at p < 0.05.

    Data availability

    The full raw dataset is provided in an additional file.
    ReferencesBarnes, D. C., Hofacer, R. D., Zaman, A. R., Rennaker, R. L. & Wilson, D. A. Olfactory perceptual stability and discrimination. Nat. Neurosci. 11, 1378–1380 (2008).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Lovitz, A. M., Sloan, A. M., Rennaker, R. L. & Wilson, D. A. Complex mixture discrimination and the role of contaminants. Chem. Senses. 37, 533–540 (2012).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Asu Schroer, S. Jakob von Uexküll: the concept of Umwelt and its potentials for an anthropology beyond the human. Ethnos 86, 132–152 (2021).Article 

    Google Scholar 
    Reinert, J. K. & Fukunaga, I. The facets of olfactory learning. Curr. Opin. Neurobiol. 76, 102623 (2022).Article 
    PubMed 

    Google Scholar 
    Johansson, B. G. & Jones, T. M. The role of chemical communication in mate choice. Biol. Rev. Camb. Philos. Soc. 82, 265–289 (2007).Article 
    PubMed 

    Google Scholar 
    Dulac, C., O’Connell, L. A. & Wu, Z. Neural control of maternal and paternal behaviors. Science 345, 765–770 (2014).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Li, Q. & Liberles, S. D. Aversion and attraction through olfaction. Curr. Biol. 25, R120–R129 (2015).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Blomkvist, A. & Hofer, M. Olfactory impairment and close social relationships. A narrative review. Chem. Senses. 46, 1–10 (2021).Article 

    Google Scholar 
    Raithel, C. U. & Gottfried, J. A. Using your nose to find your way: ethological comparisons between human and non-human species. Neurosci. Biobehav Rev. 128, 766–779 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Tabuchi, M. et al. Pheromone responsiveness threshold depends on Temporal integration by antennal lobe projection neurons. Proc. Natl. Acad. Sci. U S A. 110, 15455–15460 (2013).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kaissling, K. E. Pheromone reception in insects: The Example of Silk Moths in Neurobiology of Chemical Communication, (ed. Mucignat-Caretta, C.) Chap. 4 (CRC Press/Taylor & Francis, (2014).Gottlieb, G. Ontogenesis of sensory function in birds and mammals in The Biopsychology of Development (eds Tobach, E., Aronson, L. & Shaw, E.) 67–128 (New York Academic, (1971).Schaal, B., Orgeur, P. & Rognon, C. Odor sensing in the human fetus: anatomical functional and chemoecological bases in Fetal development: a psychobiological perspective (ed. Lecanuet, J-P, Fifer, W.P., Krasnegor, N.A., & Smotherman, edW.P.) 205–237 (Lawrence Erlbaum Hillsdale, 1995).Sullivan, R. M. Attachment figure’s regulation of infant brain and behavior. Psychodyn Psychiatry. 45, 475–498 (2017).Article 
    PubMed 

    Google Scholar 
    Blass, E. M., Teicher, M. H., Cramer, C. P., Bruno, J. P. & Hall, W. G. Olfactory, thermal, and tactile controls of suckling in preaudial and previsual rats. J. Comp. Physiol. Psychol. 91, 1248–1260 (1977).Article 

    Google Scholar 
    Pedersen, P. E. & Blass, E. M. Prenatal and postnatal determinants of the 1st suckling episode in albino rats. Dev. Psychobiol. 15, 349–355 (1982).Article 
    PubMed 

    Google Scholar 
    Johnson, M. H., Bolhuis, J. J. & Horn, G. Interaction between acquired preferences and developing predispositions during imprinting. Anim. Behav. 33, 1000–1006 (1985).Article 

    Google Scholar 
    Coureaud, G. Olfactory regulation of sucking in newborn rabbit: Ethological and chemical characterization of a pheromonal signal. Doctoral Thesis, University Paris 13 (2001).Schaal, B. et al. Chemical and behavioural characterization of the rabbit mammary pheromone. Nature 424, 68–72 (2003).Article 
    ADS 
    PubMed 

    Google Scholar 
    Coureaud, G. et al. A pheromone to behave, a pheromone to learn: the rabbit mammary pheromone. J. Comp. Physiol. A. 196, 779–790 (2010).Article 

    Google Scholar 
    Robert-Hazotte, A. et al. Nasal mucus glutathione transferase activity and impact on olfactory perception and neonatal behavior. Sci. Rep. 9, 3104 (2019).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hanser, H. I. et al. Odorant-odorant metabolic interaction, a novel actor in olfactory perception and behavioral responsiveness. Sci. Rep. 7, 10219 (2017).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Schneider, N. Y., Piccin, C., Datiche, F. & Coureaud, G. Spontaneous brain processing of the mammary pheromone in rabbit neonates prior to milk intake. Behav. Brain Res. 313, 191–200 (2016).Article 
    PubMed 

    Google Scholar 
    Coureaud, G. et al. A pheromone that rapidly promotes learning in the newborn. Curr. Biol. 16, 1956–1961 (2006).Article 
    PubMed 

    Google Scholar 
    Coureaud, G., Languille, S., Schaal, B. & Hars, B. Pheromone-induced olfactory memory in newborn rabbits: involvement of consolidation and reconsolidation processes. Learn. Mem. 16, 470–473 (2009).Article 
    PubMed 

    Google Scholar 
    Coureaud, G., Thomas-Danguin, T., Wilson, D. A. & Ferreira, G. Neonatal representation of odour objects: distinct memories of the whole and its parts. Proc. R Soc. B. 281, 20133319 (2014).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Coureaud, G., Colombel, N., Duchamp-Viret, P. & Ferreira, G. Higher-order trace conditioning in newborn rabbits. Learn. Mem. 29, 349–354 (2022).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Schneider, N. Y. et al. Brain processing of a configural vs elemental odor mixture in the newborn rabbit. Brain Struct. Funct. 221, 2527–2539 (2016).Article 
    PubMed 

    Google Scholar 
    Duchamp-Viret, P., Boyer, J., La Villa, F. & Coureaud, G. Brief olfactory learning drives perceptive sensitivity in newborn rabbits: new insights in peripheral processing of odor mixtures and induction. Physiol. Behav. 229, 113217 (2021).Article 
    PubMed 

    Google Scholar 
    Wysocki, C. J., Dorries, K. M. & Beauchamp, G. K. Ability to perceive Androstenone can be acquired by ostensibly anosmic people. Proc. Natl. Acad. Sci. U S A. 86, 7976–7978 (1989).Wang, H. W., Wysocki, C. J. & Gold, G. H. Induction of olfactory receptor sensitivity in mice. Science 260, 998–1000 (1993).Article 
    ADS 
    PubMed 

    Google Scholar 
    Wang, L., Chen, L. & Jacob, T. Evidence for peripheral plasticity in human odour response. J. Physiol. 554, 236–244 (2004).Article 
    PubMed 

    Google Scholar 
    Hummel, T. et al. Effects of olfactory training in patients with olfactory loss. Laryngoscope 119, 496–499 (2009).Article 
    PubMed 

    Google Scholar 
    Van der Linden, C., Jakob, S., Gupta, P., Dulac, C. & Santoro, S. W. Sex separation induces differences in the olfactory sensory receptor repertoires of male and female mice. Nat. Commun. 9, 5081 (2018).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Duchamp-Viret, P. & Bally, C. Plasticity mechanisms at the peripheral stage of the olfactory system. Psychol. Behav. Sci. Int. J. 17, 1–5 (2022).
    Google Scholar 
    Cain, W. S. & Drexler, M. Scope and evaluation of odor counteraction and masking. Ann. N Y Acad. Sci. 237, 427–439 (1974).Article 
    ADS 
    PubMed 

    Google Scholar 
    Thomas-Danguin, T. et al. The perception of odor objects in everyday life: a review on the processing of odor mixtures. Front. Psychol. 5, 504 (2014).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Coureaud, G., Langlois, D., Sicard, G. & Schaal, B. Newborn rabbit responsiveness to the mammary pheromone is concentration-dependent. Chem. Senses. 29, 341–350 (2004).Article 
    PubMed 

    Google Scholar 
    Cleland, T. A. et al. Sequential mechanisms underlying concentration invariance in biological olfaction. Front. Neuroeng. 4, 26–35 (2011).
    Google Scholar 
    DeChant, M. T., Bunker, P. C. & Hall, N. J. Stimulus control of odorant concentration: pilot study of generalization and discrimination of odor concentration in canines. Animals 28, 1–14 (2021).
    Google Scholar 
    Porter, D. B., Qu, L. P., Kahnt, T. & Gottfried, J. A. Aversive outcomes impact human olfactory discrimination learning and generalization. Behav. Neurosci. 135, 642–653 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Cleland, T. A., Narla, V. A. & Boudadi, K. Multiple learning parameters differentially regulate olfactory generalization. Behav. Neurosci. 123, 26–35 (2009).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Duchamp-Viret, P., Delaleu, J. C. & Duchamp, A. GABA(B)-mediated action in the frog olfactory bulb makes odor responses more salient. Neuroscience 97, 771–777 (2000).Article 
    PubMed 

    Google Scholar 
    Gutiérrez-Gálvez, A. & Marco, S. Study of the Coding Efficiency of Populations of Olfactory Receptor Neurons and Olfactory Glomeruli in Neuromorphic Olfaction (ed. Persaud, K.C., Marco, S., & Gutiérrez-Gálvez, A.) Chap. 2 (CRC Press/Taylor & Francis, (2013).Hall, N. J. & Wynne, C. D. L. Odor mixture training enhances dogs’ olfactory detection of Home-Made explosive precursors. Heliyon 4, e00947 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wright, G. A. & Smith, B. H. Stimuli and its influence on odour recognition. Proc. R Soc. B. 271, 147–152 (2004).Article 

    Google Scholar 
    Montigny, D., Coureaud, G. & Schaal, B. Rabbit pup response to the mammary pheromone: from automatism to prandial control. Physiol. Behav. 89, 742–749 (2006).Article 
    PubMed 

    Google Scholar 
    Caba, M., Lehman, M. N. & Caba-Flores, M. D. Food entrainment, arousal, and motivation in the neonatal rabbit pup. Front. Neurosci. 15, 1–10 (2021).Article 

    Google Scholar 
    Guttman, N. & Kalish, H. I. Discriminability and stimulus generalization. J. Exp. Psychol. 51, 79–88 (1956).Article 
    PubMed 

    Google Scholar 
    Shepard, R. N. Toward a universal law of generalization for psychological science. Science 237, 1317–1323 (1987).Article 
    ADS 
    MathSciNet 
    PubMed 

    Google Scholar 
    Duchamp-Viret, P., Duchamp, A. & Chaput, M. A. Peripheral odor coding in the rat and frog: quality and intensity specification. J. Neurosci. 20, 2383–2390 (2000).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Duchamp-Viret, P., Duchamp, A. & Sicard, G. Olfactory discrimination over a wide concentration range. Comparison of receptor cell and bulb neuron abilities. Brain Res. 517, 256–262 (1990).Article 
    PubMed 

    Google Scholar 
    Wilson, D. A. Pattern separation and completion in olfaction. Ann. N Y Acad. Sci. 1170, 306–312 (2009).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    DeChant, M. T. & Hall, N. J. Training with varying odor concentrations: implications for odor detection thresholds in canines. Anim. Cogn. 24, 889–896 (2021).Article 
    PubMed 

    Google Scholar 
    Larsson, L. & Laska, M. Ultra-high olfactory sensitivity for the human sperm-attractant aromatic aldehyde bourgeonal in CD-1 mice. Neurosci. Res. 71, 355–360 (2011).Article 
    PubMed 

    Google Scholar 
    Sarrafchi, A., Odhammer, A. M., Hernandez Salazar, L. T. & Laska, M. Olfactory sensitivity for six predator odorants in CD-1 mice, human subjects, and spider monkeys. PLoS One. 20, e80621 (2013).Article 

    Google Scholar 
    Beidler Walker, D. et al. Naturalistic quantification of canine olfactory sensitivity. Appl. Anim. Behav. Sci. 97, 241–254 (2006).Article 

    Google Scholar 
    Concha, A. R. et al. Canine olfactory thresholds to amyl acetate in a biomedical detection scenario. Front. Vet. Sci. 5, 1–11 (2019).Article 

    Google Scholar 
    Fujiwara, T. et al. Odorant concentration differentiator for intermittent olfactory signals. J. Neurosci. 34, 16581–16593 (2014).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Jinn, J., Connor, E. G. & Jacobs, L. F. How ambient environment influences olfactory orientation in search and rescue dogs. Chem. Senses. 45, 625–634 (2020).Article 
    PubMed 

    Google Scholar 
    Kokocinska-Kusiak, A. et al. Canine olfaction: physiology, behavior, and possibilities for practical applications. Animals 11, 2463 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kadakia, N. et al. Odour motion sensing enhances navigation of complex plumes. Nature 611, 754–761 (2022).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Nevitt, G. A., Dittman, A. H., Quinn, T. P. & Moody, W. J. Jr Evidence for a peripheral olfactory memory in imprinted salmon. Proc. Natl. Acad. Sci. USA. 91, 4288–4292 (1994).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Zarrow, M. X., Denenberg, V. H. & Anderson, C. O. Rabbit: frequency of suckling in the pup. Science 150, 1835–1836 (1965).Article 
    ADS 
    PubMed 

    Google Scholar 
    Coureaud, G., Fortun-Lamothe, L., Langlois, D. & Schaal, B. The reactivity of neonatal rabbits to the mammary pheromone as a probe for viability. Animal 1, 1026–1032 (2007).Article 
    PubMed 

    Google Scholar 
    Coureaud, G., Letagneaux, C., Thomas-Danguin, T. & Romagny, S. Developmental changes in elemental and configural perception of odor mixtures in young rabbits. Dev. Psychobiol. 62, 471–483 (2020).Article 
    PubMed 

    Google Scholar 
    Download referencesAcknowledgementsWe sincerely thank Priscilla Orlando, Gianni Raponi, Clarisse Auclair, Océane Meunier, Manon Dirheimer from CRNL Animal facility, Jiasmine Boyer, Florian La Villa and Thierry Thomas-Danguin for their collaboration on this study, and Kasia Pisanski and David Reby for English editing of the manuscript.This project received support from the French Agence Nationale de la Recherche (ANR) under the NEONATOLF Grant agreement (ANR-20-CE20-0019-01) to G.C., M.S.H., P.D.V. and J.M.H. The authors also acknowledge general support from the Centre National de la Recherche Scientifique (CNRS) and Institut National de Recherche pour l’Agriculture, l’Alimentation et l’Environnement (INRAE).Author informationAuthor notesGérard Coureaud and Marie-Sabelle Hjeij contributed equally to this work.Authors and AffiliationsLyon Neuroscience Research Center, CNRS UMR 5292, INSERM U 1028, Université Claude Bernard Lyon 1, Université Jean Monnet, Centre Hospitalier Le Vinatier – Bâtiment Neurocampus, 40 Avenue du Doyen Lepine, Bron, F-69500, FranceGérard Coureaud, Marie-Sabelle Hjeij, Jeanne Serrano, Marc Thévenet, Samuel Garcia & Patricia Duchamp-ViretLaboratoire Interdisciplinaire Carnot de Bourgogne ICB UMR 6303, Université Bourgogne Europe, CNRS, 7 bvd Jeanne d’Arc, Dijon, F-21000, FranceJean-Marie HeydelAuthorsGérard CoureaudView author publicationsSearch author on:PubMed Google ScholarMarie-Sabelle HjeijView author publicationsSearch author on:PubMed Google ScholarJeanne SerranoView author publicationsSearch author on:PubMed Google ScholarMarc ThévenetView author publicationsSearch author on:PubMed Google ScholarSamuel GarciaView author publicationsSearch author on:PubMed Google ScholarJean-Marie HeydelView author publicationsSearch author on:PubMed Google ScholarPatricia Duchamp-ViretView author publicationsSearch author on:PubMed Google ScholarContributionsConceptualization: G.C., P.D.V.; Methodology: G.C., P.D.V.; Investigation: G.C., M.S.H., J.S.; Technical support: M.T., S.G.; Formal analysis: G.C., M.S.H., P.D.V.; Writing – original draft: G.C., M.S.H., P.D.V.; Writing – review and editing: G.C., M.S.H., J.S., J.M.H., P.D.V.; Supervision: G.C., P.D.V; Funding acquisition: G.C., J.M.H., P.D.V.Corresponding authorCorrespondence to
    Gérard Coureaud.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 2Rights 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 articleCoureaud, G., Hjeij, MS., Serrano, J. et al. Odour learning concentration influences concentration range of conditioned response in newborn rabbits.
    Sci Rep 15, 44682 (2025). https://doi.org/10.1038/s41598-025-28326-6Download citationReceived: 03 April 2025Accepted: 10 November 2025Published: 29 December 2025Version of record: 29 December 2025DOI: https://doi.org/10.1038/s41598-025-28326-6Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative
    KeywordsNewbornPheromoneOdour learningConcentrationPerceptionRabbit More

  • in

    Satellites reveal widespread deoxygenation of large global reservoirs from 1984 to 2023

    AbstractDissolved oxygen (DO) in reservoirs regulate biodiversity, nutrient biogeochemistry, water quality, and greenhouse gas emissions. Maintaining healthy DO levels is essential for achieving United Nations Sustainable Development Goals (SDGs), particularly SDG 6 (Clean Water and Sanitation) and SDG 14 (Life Below Water). However, our full understanding of long-term DO dynamics of global reservoirs remains unknown, due to limited observations. Here, we develop a satellite-based machine learning model to reveal DO dynamics of major reservoirs (area > 100 km2) at the global scale. We first based on continuous DO in-situ records (containing ~ 32,065 samples) to comprehensively evaluate the performance of estimating DO using three widely-used machine learning methods (e.g., Random Forest, RF; eXtreme Gradient Boosting, XGBoost; and Support Vector Regression, SVR). The RF outperforms other methods and can reliably estimate DO with R2 = 0.73 and RMSE = 1.23 mg/L in testing set. Our results demonstrate that global reservoirs show widespread deoxygenation (74%, 264 out of 357) from 1984 to 2023, with an average DO decreasing rate of 0.13 mg/L per decade, which is faster than that observed in the lakes, oceans, and rivers. Reservoir DO exhibits pronounced spatial heterogeneity, with DO in cold northern systems is approximately 1.5 times that of in tropical and regions, reflecting latitudinal, climatic, and continental contrasts. These rapidly declining DO are mainly controlled by climate changes (contributing ~ 46%), human perturbations (contributing ~ 31%, through land use change and nutrient inputs), and the biogeochemical processes (contributing ~ 23%, through primary production and turbidity), as quantified by a state-of-the-art machine learning-based attribution analysis (SHapley Additive exPlanations, SHAP). Our study presents a practical method for spatiotemporal reconstruction of global reservoir DO dynamics using remote sensing and contributes to better understanding of driving factors behind DO changes in major reservoirs.

    Similar content being viewed by others

    Dynamic monitoring of ecological security patterns in arid zone oases: a remote sensing-based ecological index evolution analysis

    Article
    Open access
    10 December 2025

    Dynamics of oxygen evolution in a thermally stratified reservoir under climate warming

    Article
    Open access
    18 November 2025

    Advanced generalized machine learning models for predicting hydrogen–brine interfacial tension in underground hydrogen storage systems

    Article
    Open access
    30 May 2025

    IntroductionDissolved oxygen (DO) is a fundamental indicator of water quality and ecosystem health in aquatic systems1,2,3. It regulates the survival, growth, and distribution of fish, invertebrates, and microorganisms, while also controlling key biogeochemical processes such as nutrient cycling, carbon turnover, and greenhouse gas emissions4. Deoxygenation—manifesting as hypoxia (DO < 2 mg/L) or anoxia (DO < 0.5 mg/L)—has severe ecological and societal consequences, including fish kills, harmful algal blooms, loss of biodiversity, and deterioration of drinking water quality5,6,7. These processes are especially pronounced in reservoirs, where water-column stratification and anthropogenic regulation often exacerbate oxygen depletion compared with many natural lakes or rivers, although some deep natural lakes also develop strong stratification and some shallow lakes experience substantial anthropogenic impacts3,8,9,10.Reservoirs are highly sensitive to deoxygenation due to their dual roles as engineered systems and aquatic ecosystems11,12. Thermal and chemical stratification restrict vertical mixing, confining oxygenated waters to the surface and promoting hypoxia or anoxia in bottom layers13. This process not only reduces habitat availability for aquatic life but also enhances the release of phosphorus, nitrogen, and reduced metals from sediments, fueling eutrophication and impairing water supply5. Moreover, anthropogenic drivers—including damming, altered hydrological regimes, land-use change, and nutrient loading—intensify oxygen depletion and amplify climate-induced warming effects14,15. Collectively, these processes make reservoirs hotspots of rapid and widespread deoxygenation, with cascading consequences for aquatic biodiversity, nutrient cycling, and human health2,3,7.Despite its importance, long-term monitoring of DO in reservoirs remains challenging. Traditional in-situ measurements provide accurate but highly localized data16,17, often failing to capture the strong spatial and temporal heterogeneity of oxygen dynamics in large and complex reservoirs6,9,18,19. The labor-intensive nature of field sampling limits the temporal coverage needed to assess seasonal and interannual trends, while sparse monitoring networks are insufficient to characterize basin-wide deoxygenation patterns18. These constraints underscore the need for innovative monitoring strategies capable of reconstructing historical DO variability across broad spatial and temporal scales20,21,22.The integration of remote sensing and machine learning provides significant advantages for analyzing deoxygenation in large spatio-temporal scales. Satellite data offers broad-scale, high-frequency coverage, overcoming the limitations of traditional point sampling methods23. Specifically, Landsat archives provide spectral data with a 16-day revisit period and 30-m spatial resolution, containing information on water constituents like chlorophyll-a, suspended sediment, and dissolved organic matter, all of which are closely related to DO production and depletion24. Furthermore, machine learning algorithms are crucial for interpreting complex, multi-dimensional remote sensing data, enabling the identification of intricate, non-linear relationships between spectral signatures and DO levels. Recent studies have successfully used machine learning to retrieve DO concentrations from satellite data and even predict deoxygenation events6,7,9,18,19. This allows us to reconstruct historical DO data over the past four decades, facilitating in-depth analysis of long-term trends and patterns. Thus, the integration of remote sensing and machine learning offers a powerful tool for understanding deoxygenation of reservoirs globally.In this study, we present a generalized global reservoir DO retrieval method utilizing four decades of Landsat data and multiple machine learning algorithms to reconstruct the temporal dynamics of DO in 357 major reservoirs (area > 100 km2) globally. We first compile a comprehensive DO dataset covering the globe, ensuring sufficient data for the development of robust machine learning models. We then use our DO retrieval model to: (1) characterize multi-decadal spatiotemporal patterns of reservoir deoxygenation, (2) quantify the four-decade trends in reservoir deoxygenation, and (3) disentangle the relative contributions of climate warming and anthropogenic regulation to observed DO changes. This study introduces a practical methodology for large-scale spatiotemporal reconstruction of DO using remote sensing and machine learning, enhancing our understanding of human impacts and climate change on DO dynamics in global reservoirs.Data and methodsIn-situ DO data and matchupField DO observations are downloaded from several public sources including the Global River Water Quality Archive (GRQA)17 and the Global Freshwater Quality dataset (accessible at https://gemstat.bafg.de/). We first filter the in-situ observations to the period 1984–2023, corresponding to the temporal coverage of the satellite record. We then implement additional manual quality control (e.g., applying a strict spatio-temporal matching window and removing low-quality pixels) to identify suitable satellite-field data matchups for algorithm development (Fig. 1). To accurately identify water surfaces, we employ the widely used Dynamic Surface Water Extent (DSWE) method25. We adopt a widely used cross-sensor harmonization approach to ensure consistency in sensor characteristics and minimize discrepancies from multi-sensor integration, including Landsat ETM, ETM+, and OLI. Specifically, we apply cross-sensor calibration following established procedures23, which have been demonstrated to reduce systematic biases among Landsat sensors. We implement rigorous temporal matching criteria, restricting the time difference between field measurements and satellite overpasses to a strict ± 6 h window. Spatial data processing involved extracting a 3 × 3 pixel window surrounding each sampling location to mitigate adjacency effects. This is followed by comprehensive quality screening using QA_PIXEL band bitmask technology to systematically exclude pixels contaminated by clouds, snow, or cloud shadows. Only windows retaining over 50% valid pixels, defined as pixels free from cloud, cloud shadow, and snow contamination according to the QA_PIXEL bitmask, with a spatial coefficient of variation below 0.15 (calculated within the 3 × 3 pixel window) underwent further analysis. The final reflectance values for all six Landsat spectral bands (blue, green, red, near-infrared, SWIR1, and SWIR2) calculated as the mean of these quality-controlled pixels. Following these procedures, we successfully match a total of 32,065 DO measurements globally, including 1,732 matchups for the studied reservoirs. Fig. 1The spatial mapping of global DO concentration measurements. Field DO observations are downloaded from several public sources including the Global River Water Quality Archive (GRQA)17 and the Global Freshwater Quality dataset (accessible at https://gemstat.bafg.de/).Full size imageSatellite data and pre-processingThe USGS Landsat archive offers valuable long-term satellite data (1984–2023), enabling scientific research on historical water quality parameter retrieval. We first select major reservoirs (area > 100 km2) based on boundaries provided by the Global Surface Water (GSW)26. Given that reliable global reservoir volume data are still limited and often inconsistent due to variations in water depth and bathymetric uncertainty26, we used surface area as a practical and consistent criterion for reservoir selection. Reservoir surface area and volume are strongly correlated globally (R2 = 0.76; Supplementary Fig. 1), supporting the use of area (> 100 km²) as a robust proxy for identifying major reservoirs in this study. We define permanent water as pixels with ≥ 75% water occurrence frequency, based on global imagery from the GSW dataset (30 × 30 m resolution) and considering seasonal water body variations18. We acquire Landsat-5, Landsat-7, and Landsat-8 images from the United States Geological Survey (USGS) archive and process them using ACOLITE27, a widely used atmospheric correction method. We then use a widely used method to maintain sensor consistency across the four-decade study period28. Cloud-contaminated scenes are filtered using Landsat’s internal metadata, retaining only images with less than 30% cloud cover. Additionally, rigorous pixel-level quality control was applied using the QA_PIXEL band bitmask to eliminate unreliable observations, including those affected by clouds, snow, or cloud shadows. Water pixels are identified using the USGS-developed DSWE algorithm25. Finally, we obtain 159,969 processed images for DO retrieval (Fig. 1).Machine learning retrieval model developmentTo predict DO, we select three representative machine learning algorithms: Random Forest (RF), eXtreme Gradient Boosting (XGBoost), and Support Vector Regression (SVR). Model development is conducted using the scikit-learn package (version 1.3.1) in Python 3.8. Random samples from the integrated dataset are partitioned into training (70%) and testing (30%) sets. Hyperparameters (e.g., number of trees, maximum depth, learning rate, and kernel parameters) are optimized using a grid search strategy coupled with 10-fold cross-validation. We incorporate remote sensing reflectance (e.g., visible bands and near-infrared band), hue angle, longitude, latitude, and DEM information into the model following a recent study18. SWIR bands are excluded because strong water absorption in the SWIR (especially > ~ 1,300 nm) yields negligible water-leaving reflectance and low signal/noise ratio for most inland waters29. Hue angle serves as a robust proxy for water color, and is widely used to assess water quality in various water bodies18. More details of the hue angle can be found in Hu et al.18. Feature importance is quantified via SHapley Additive exPlanations (SHAP) values. Details of the optimized hyperparameters are provided in Table 1. Each of these three algorithms offers distinct advantages: RF is an ensemble learning method that aggregates predictions from multiple independent decision trees. It is well-regarded for its robustness and is widely applied in water quality research; XGBoost is a powerful boosting ensemble algorithm built on gradient-boosted decision trees. It incorporates regularization techniques to control model complexity, prevent overfitting, and enhance computational efficiency. Previous studies have also highlighted its strong predictive capabilities and superior performance in water quality parameter estimation. Support Vector Regression (SVR), grounded in structural risk minimization theory, excels at handling small sample sizes and nonlinear problems, commonly applied in fields like water quality retrieval by satellites.Table 1 Training hyperparameters of machine learning models.Full size tableSpatiotemporal analysisTo minimize the influence of ice cover on DO estimates, we focus on satellite imagery acquired during the ice-free periods: July-September in the Northern Hemisphere and January-March in the Southern Hemisphere7,30. The average DO concentration during these months is used to represent each reservoir and year. To analyze long-term trends in annual DO, we use the non-parametric Mann-Kendall test, setting the significance level (α) to 0.05. We also investigate the relationships between DO and other environmental variables using Pearson correlation. The statistical significance of these correlations, as well as differences between group means, is assessed using a t-test with a significance threshold of p < 0.05. While Pearson’s correlation coefficient quantifies the strength and direction of the linear relationship between two continuous variables, the t-test determines whether the observed relationship or difference is statistically significant, taking into account the sample size and variability.Quantification of contributions of DO change driversThe pervasive decline in reservoirs DO concentrations stems from a complex interplay of environmental factors, anthropogenic perturbations, and biogeochemical processes18,20,21. Disentangling the roles of these various drivers is paramount for understanding reservoir ecosystem development and safeguarding aquatic biomass health21. To precisely quantify their individual contributions, we employ a machine learning attribution method. This comprehensive analysis encompassed environmental factors (e.g., temperature, precipitation, wind, and evaporation), human activities (e.g., populations, urban area, and cropland area), and other key biogeochemical processes (e.g., turbidity and primary production). To improve the physical representation of temperature effects on DO solubility, we reprocess Landsat Level-2 Surface Temperature (LST) data across all studied reservoirs during ice-free months and use these observations to calibrate ERA5 air temperature via regression correction (Supplementary Fig. 2). The corrected temperatures are subsequently used for model development and driver attribution. Our data sources include the ECMWF Reanalysis v5 (ERA5) Land product for environmental factors, the MODIS Land Cover Type Product (https://lpdaac.usgs.gov/products/mcd12q1v006/) for urban and cropland areas, and LandScan (https://landscan.ornl.gov) for population data. Reservoir turbidity is precisely calculated using the Normalized Difference Turbidity Index (NDTI), which is a widely used proxy for aquatic turbidity levels.Model performance evaluationModel performance is comprehensively evaluated using two key metrics: the coefficient of determination (R2) and the root mean squared error (RMSE). R2 quantifies the proportion of variance in the dependent variable that is predictable from the independent variables, with values ranging from 0 to 1. Higher R2 values indicate a better model fit to the observed data. Conversely, RMSE measures the average magnitude of the errors between predicted and observed values. Lower RMSE values denote higher model prediction accuracy. The formulas for calculating these evaluation metrics are as follows:$$:{R}^{2}=1-sum:_{i=1}^{N}frac{{({X}_{i}-{Y}_{i})}^{2}}{{({X}_{i}-stackrel{-}{Y})}^{2}}$$
    (1)
    $$:RMSE=sqrt{frac{1}{N}sum:_{i=1}^{N}{({X}_{i}-{Y}_{i})}^{2}}$$
    (2)
    where (:{X}_{i}) represents the in-situ measured data, and (:{Y}_{i}) represents the retrieved data.ResultsCharacteristics of observed DO dataThe global assessment of reservoir dissolved oxygen (DO) concentrations reveals pronounced intercontinental disparities (Fig. 2). Europe exhibits a notably high mean DO level of 9.8 ± 1.3 mg/L, supported by a robust sample size (N = 11,525). Africa records the highest mean concentration (10.8 ± 2.1 mg/L), which may reflect relatively undisturbed, pristine aquatic environments; however, this interpretation warrants caution given the smaller sample size (N = 903). In contrast, the moderate DO levels observed in Asia (8.2 ± 1.3 mg/L; N = 13,300) and North America (7.9 ± 1.2 mg/L; N = 2,158) likely result from the complex interplay between extensive anthropogenic pressures and ongoing mitigation measures. More concerning patterns emerge in South America and Oceania, where the comparatively low mean DO concentrations (6.7 ± 0.8 mg/L and 6.3 ± 1.1 mg/L, respectively) suggest heightened anthropogenic stress or distinct regional environmental constraints, underscoring the need for further investigation and targeted management strategies.Fig. 2The continental pattern of in-situ DO data. On boxes, the center line shows the median values, the hollow circle shows the mean values, the whiskers denote the full range (min and max), and box limits indicate the 25th and 75th percentiles.Full size imageModel performanceOur comparison of three machine learning methods reveals marked differences in their ability to accurately estimate reservoirs DO concentration (Fig. 3). The RF model demonstrates the highest performance, achieving an R2 of 0.77 and a RMSE of 1.12 mg/L in the training set, and an R2 of 0.73 with a RMSE of 1.23 mg/L in the testing set. The RF also shows a relatively tight clustering of predicted versus observed DO concentrations closely aligned with the 1:1 line, indicating a strong capability in capturing the complex nonlinear relationship between satellite data and DO. In contrast, the XGBoost model shows a moderate performance with an R2 of 0.58 and an RMSE of 1.59 mg/L in testing set, with some uncertainties at higher DO concentrations (e.g., higher than 15 mg/L), and some banding at lower values (e.g., lower than 5 mg/L). The SVR model exhibits lower accuracy, with an R2 of 0.45 and an RMSE of 1.93 mg/L in testing set, showing a considerable spread of data points and a tendency to under-predict higher DO values. With 10-fold cross-validation, the RF achieves robust DO estimation (R2 = 0.72, RMSE = 1.27 mg/L), representing a significant improvement over other methods, which have R2 values ranging from 0.42 to 0.56 and RMSE values from 1.63 to 1.91 mg/L. Based on these comparative metrics, the RF clearly stands out as the most effective algorithm among the tested methods for reliably estimating reservoir DO concentrations in this study (Fig. 3). To evaluate the model performance at the study sites, we compared satellite-derived and measured DO concentrations across 326 reservoirs with available in situ measurements (total 1,732 matchups). The RF model achieve an R2 of 0.69 and an RMSE of 1.33 mg/L (Fig. 4).Fig. 3The comparison of model performance of three machine learning models, e.g., Random Forest (RF), eXtreme Gradient Boosting (XGBoost), and Support Vector Regression (SVR). The RF model outperform other models with R2 = 0.73 and RMSE = 1.23 mg/L in testing set.Full size imageFig. 4The model performance of RF across 357 studied reservoirs.Full size imageThe superior performance of the RF model may be attributed to its ensemble structure, which effectively captures complex nonlinear relationships and interactions among multiple predictors while being less sensitive to parameter tuning31. By contrast, SVR can struggle with highly nonlinear dynamics in large, heterogeneous datasets, often leading to underestimation of high DO values32. XGBoost, although powerful, may require more extensive parameter optimization and larger training datasets to achieve comparable performance, and can be prone to overfitting when the number of features is relatively small33. These characteristics likely explain the observed differences in model accuracy in our study context.The spatial pattern of global major reservoirs DO concentrationsOur four-decade global assessment of DO dynamics across global 357 major reservoirs reveal pronounced spatial heterogeneity (Fig. 5). From 1984 to 2023, the global mean annual surface DO concentration is 8.4 ± 1.1 mg/L (mean ± SD). A distinct hemispheric contrast emerges: reservoirs in the Northern Hemisphere maintain higher DO levels (9.8 ± 0.9 mg/L) compared to those in the Southern Hemisphere (7.7 ± 1.4 mg/L) (Fig. 6). These concentrations generally remain sufficient to support diverse fish assemblages and avoid the stress associated with extreme oxygen depletion, positioning many reservoirs as critical ecological benchmarks for sustaining aquatic biodiversity.Fig. 5The spatial mapping of DO in global reservoirs. A striking spatial contrast exists in DO concentrations across global reservoirs: higher concentrations are prevalent in the Northern Hemisphere, while lower concentrations dominate the Southern Hemisphere.Full size imageFig. 6Spatio-temporal distribution of DO concentrations across different latitudinal gradients and climatic zones. (a) Spatial distribution of DO concentrations. (b) Long-term DO concentrations across different continents. North American and European reservoirs show significantly higher DO levels than those in Africa, Asia, and South America. The red line represents mean value of six continents. In the boxplots, the horizontal line within each box represents the median value, the width of the box indicates the interquartile range (IQR), and the whiskers extend to a maximum of 1.5 times the IQR from the box edges.Full size imageLatitudinal gradients further highlight this heterogeneity (Fig. 5). Polar reservoirs (latitude > 66.56°) exhibit the highest DO concentrations (10.8 ± 0.5 mg/L), whereas equatorial reservoirs (5°S to 5°N) display the lowest (6.9 ± 0.8 mg/L). The extremes of this spectrum range from a minimum of 5.2 mg/L near the equator to a maximum of 11.5 mg/L in Arctic systems. The elevated DO levels in cold-region reservoirs exemplify pristine oxygenation, largely sustained by oligotrophic conditions, cold montane hydrology, and minimal anthropogenic disturbance. When classified by climatic zones, the North Frigid (10.8 ± 0.5 mg/L) and North Temperate zones (9.6 ± 0.7 mg/L) consistently maintain higher DO than the Tropical (6.9 ± 0.6 mg/L) and South Temperate zones (8.7 ± 1.1 mg/L, median 9.1 mg/L). Variability, expressed as the standard deviation of multi-decadal DO, is also higher in the North Frigid (0.6 ± 0.1 mg/L) and Tropical zones (0.6 ± 0.3 mg/L) compared with the more stable North Temperate (0.5 ± 0.1 mg/L) and South Temperate zones (0.4 ± 0.1 mg/L).At the continental scale, North American (9.2 ± 0.8 mg/L) and European (7.5 ± 0.8 mg/L) reservoirs show significantly higher DO levels than those in Africa, Asia, and South America, where concentrations typically range between 7 and 8 mg/L (Fig. 6). However, localized low-oxygen anomalies occur in parts of Africa, Asia, and South America, with DO values falling well below the continental mean (8.4 ± 1.3 mg/L). This north–south disparity reflects interacting climatic and ecological drivers: northern reservoirs benefit from lower thermal regimes that enhance oxygen solubility, extensive riparian forest cover that mitigates nutrient loading, and wind-driven mixing that promotes vertical oxygenation, whereas equatorial and subtropical systems remain more vulnerable to thermal stratification and eutrophication2,34. These findings not only highlight the global heterogeneity of reservoir oxygenation but also signal potential climate vulnerabilities. Especially, we find that reservoirs in North and South America, already exhibiting oxygen stress with some outliers significantly below the median, are particularly at risk of further deoxygenation (Fig. 6).The temporal pattern of DO concentrationsWe find that reservoirs worldwide show a decreasing trend in surface DO, with a mean rate of − 0.13 mg/L per decade (Fig. 7). This trend is not uniform across all reservoirs; globally, 74% of the studied reservoirs experienced a decrease in surface DO, averaging − 0.07 mg/L per decade, while 26% show an increase at a rate of 0.07 mg/L per decade. Reservoirs with increasing DO are predominantly located in tropical regions. This decline in reservoir surface DO is widespread on a continental level, with all six continents studied showing decreases over the past four decades: Africa (− 0.07 mg/L), Asia (− 0.06 mg/L), Europe (− 0.12 mg/L), North America (− 0.10 mg/L), Oceania (− 0.05 mg/L), and South America (− 0.05 mg/L, Fig. 8).Fig. 7Four decadal DO changes in global reservoirs. The DO trend is derived using a two-sided Mann–Kendall test at a 95% confidence level. Our results demonstrate that global reservoirs show widespread deoxygenation (74%, 264 out of 357) from 1984 to 2023, with an average DO decreasing rate of 0.13 mg/L per decade.Full size imageFig. 8Four decadal DO changes in global reservoirs. The DO trend is derived using a two-sided Mann–Kendall test at a 95% confidence level. Our results demonstrate that global reservoirs show widespread deoxygenation (74%, 264 out of 357) from 1984 to 2023, with an average DO decreasing rate of 0.13 mg/L per decade. In the boxplots, the horizontal line within each box represents the median value, the width of the box indicates the interquartile range (IQR), and the whiskers extend to a maximum of 1.5 times the IQR from the box edges.Full size imageSeasonal surface DO of the studied reservoirs show greater variability compared to their interannual variations (Fig. 9). Seasonal surface DO fluctuations are greater in the Northern Hemisphere (1.8 ± 0.3 mg/L) and smaller in the Southern Hemisphere (1.1 ± 0.2 mg/L). Seasonal variations in DO concentrations display divergent patterns between the Southern Hemisphere and the Northern Hemisphere (Fig. 9). In the Southern Hemisphere, mean DO levels peak around July (8.1 mg/L) and reach their minimum in March (7.0 mg/L). Conversely, in the Northern Hemisphere, surface DO levels peak around January (11.3 mg/L) and are lowest around July (7.4 mg/L). Furthermore, the monthly averaged DO in the Southern and Northern Hemispheres shows contrasting trends: a convex pattern predominates in the Southern Hemisphere, while a combination of convex and concave patterns is observed in the Northern Hemisphere (Fig. 9). This pattern aligns with typical temperate reservoir dynamics where colder months favor higher oxygen solubility and reduced biological oxygen demand, while warmer periods experience both decreased solubility and increased respiratory consumption.Fig. 9Seasonal patterns of global reservoir DO concentrations. Specifically, a clear divergence exists between the hemispheres: the Southern Hemisphere displays a convex DO pattern, in contrast to the Northern Hemisphere, which exhibits a combination of convex and concave patterns.Full size imageDiscussionThe controls of widespread decreases of DO concentrationsOur results reveal that the rates of change in reservoir surface DO is faster than those observed in the lakes7 (− 0.08 mg/L per decade), oceans20 (− 0.02 mg/L per decade) and in rivers3 (− 0.04 mg/L per decade) over a similar time period. This highlights that reservoirs are particularly vulnerable to deoxygenation, likely due to the combined effects of anthropogenic regulation and climate forcing. SHAP analysis identifies temperature (19.4%) as the predominant driver of DO decline, a finding consistent with global warming impacts7 (Fig. 10). Although oxygen solubility is temperature-dependent34, our analysis specifically focuses on summer (ice-free) months when intra-hemispheric temperature variability is relatively small. Moreover, by explicitly quantifying temperature’s contribution through SHAP analysis, we effectively account for thermal effects on DO dynamics without converting DO concentrations to percent saturation. This approach ensures that the observed spatial and temporal patterns reflect genuine biogeochemical and anthropogenic influences rather than artifacts of solubility normalization. Rising water temperatures directly reduce oxygen solubility while enhancing microbial respiration, creating a synergistic mechanism for oxygen depletion18. Cropland (13.9%) emerges as the secondary driver, largely through agricultural runoff (fertilizers, pesticides), which triggers eutrophication and promotes algal blooms, whose decomposition by heterotrophic bacteria further depletes DO via biochemical oxygen demand2,34 (Fig. 10). However, we acknowledge that reservoirs artificially regulated by dams may be less affected by agricultural fertilization activities compared to natural lakes, yet previous studies have demonstrated that nutrient enrichment can exert long-term controls on DO dynamics through eutrophication processes35. Primary production (12.7%) affects DO by increasing oxygen through photosynthesis, but when nutrient enrichment promotes algal biomass accumulation, the later breakdown of this biomass intensifies microbial oxygen consumption and reduces5,21,36. Together, these results reveal a dual-threat framework, which climate-driven thermal stress and nutrient pollution act in concert to amplify hypoxia risk in aquatic ecosystems (Fig. 10).Fig. 10Random Forest attribution analysis identifies the contribution of various drivers to the model’s predictions. Horizontal bars visualize the relative importance of each driver, showing its influence on the model output. Longer bars indicate greater importance, highlighting the key factors driving the model’s predictions.Full size imageBeyond primary drivers, secondary mechanisms exert nuanced but critical influences on deoxygenation dynamics. Wind speed (12.3%) demonstrates dual modulation, e.g., wind-driven mixing enhances surface aeration, whereas prolonged calm periods promote stratification, isolating hypoxic waters in benthic zones7 (Fig. 10). Turbidity (10.7%), largely driven by sediment influx, attenuates photic zone depth, concurrently inhibiting photosynthetic O2 production and stimulating organic matter mineralization5,21. Anthropogenic factors are equally pivotal, e.g., urban land cover (8.9%) and population density (8.2%) drive hypoxia through wastewater effluents and nutrient leaching from impervious surfaces21 (Fig. 10). Precipitation (5.4%) exhibits paradoxical effects, which dilutional oxygenation versus pulse-loaded nutrient surges34. Ecological shifts captured by NDVI change rate (5.0%) modulate nutrient retention, while evaporation (3.4%) concentrates dissolved constituents, exacerbating oxygen deficits through volumetric compression7. Collectively, these interdependencies generate nonlinear feedbacks that govern the basin’s spatiotemporal deoxygenation gradients.Beyond above climatic and watershed drivers, several reservoir-specific factors also play critical roles in shaping DO dynamics. Unlike natural lakes, reservoirs are subject to direct human regulation, including dam release operations, seasonal water storage, and hydropower generation, all of which can alter vertical mixing and stratification regimes35. For instance, selective withdrawal during dam releases often exports hypolimnetic water with low oxygen concentrations, further exacerbating downstream hypoxia3. Moreover, reservoir morphology—such as mean depth, bathymetry, and shoreline complexity—can interact with meteorological processes to influence oxygen dynamics37. Shallow or morphologically complex reservoirs tend to experience stronger wind-driven mixing and shorter stratification periods, whereas deeper reservoirs with smoother bathymetry are more prone to persistent hypolimnetic oxygen depletion. Although these morphological and operational factors are challenging to quantify globally, their potential impacts on reservoir deoxygenation warrant further investigation in future research.Model advantage and limitationIn this study, we develop a generalized method to robustly reconstruct the spatio-temporal patterns of deoxygenation across major reservoirs globally. Our machine learning framework further enhances the ability to downscale and harmonize satellite-derived water quality metrics with historical DO measurements, enabling a multi-decadal reconstruction of deoxygenation trends with R2 is 0.73 and RMSE is 1.23 mg/L in testing set (Figs. 3 and 4). This approach not only captures interannual variability driven by climate oscillations but also identifies hotspots of hypoxia linked to anthropogenic pressures such as agricultural runoff and urbanization.However, while satellite remote sensing provides invaluable insights, several limitations must be acknowledged. First, cloud cover and atmospheric interference can lead to data gaps, particularly in temperate regions with frequent cloudiness6,38. Second, current satellite sensors are constrained by their revisit cycles (e.g., the temporal resolution of Landsat at 16 days), making it challenging to achieve daily-scale monitoring of rapidly changing oxygen dynamics, such as those driven by storm events or sudden algal blooms3. Despite these constraints, our methodology optimizes the use of available satellite data by leveraging statistical gap-filling techniques and multi-sensor fusion to improve temporal resolution. Future advancements in hyperspectral satellites (e.g., NASA’s PACE mission) and geostationary water quality monitoring could further enhance our ability to track reservoir deoxygenation at finer spatio-temporal scales39. By combining satellite observations with emerging autonomous in situ sensors (e.g., gliders, buoys), we can move toward a more comprehensive understanding of deoxygenation mechanisms in inland waters28.In addition, although our DO retrieval model achieves reasonable predictive results (R² = 0.73, RMSE = 1.23 mg/L), we acknowledge that this level of uncertainty should be carefully considered. Specifically, an error of ~ 1.23 mg/L represents approximately 15% of the global mean DO concentration (8.4 ± 1.1 mg/L), which may affect the precise quantification of DO at individual reservoirs, especially in aquatic systems with relatively low DO concentrations where small biases could alter ecological thresholds such as hypoxia (DO < 2 mg/L). Overall, our study demonstrates that remote sensing, despite its limitations, is indispensable for large-scale, long-term reservoir deoxygenation assessments, providing critical data to inform water resource management and climate adaptation strategies.While our study focuses on reservoirs larger than 100 km2, the methodology could, in principle, be extended to smaller reservoirs or other types of inland waters. However, smaller water bodies often exhibit more rapid hydrodynamic responses, higher variability in thermal stratification, and distinct catchment influences11, which may require additional calibration of the machine learning model. Future work could leverage higher-resolution satellite imagery, autonomous in situ sensors, and targeted field campaigns to adapt and validate the model for these smaller systems. Such efforts would expand the applicability of our framework, enabling comprehensive assessments of oxygen dynamics across a wider range of inland waters.Implication for reservoirs management and the sustainable development goalsOur reconstruction of four decades of surface DO dynamics in global reservoirs provides critical insights into both ecological processes and reservoirs management strategies11. Reservoirs serve as vital water sources for drinking, irrigation, and hydropower production, while simultaneously sustaining diverse aquatic ecosystems37. Progressive deoxygenation directly threatens these functions by reducing habitat suitability for fish, altering food-web dynamics, and accelerating biodiversity loss40,41. Low oxygen conditions also facilitate the release of nutrients and metals from sediments, thereby promoting harmful algal blooms and compromising drinking water quality1,42. These effects are of particular concern in densely populated regions where reservoirs are the primary sources of freshwater supply43. Despite their relatively high present-day DO concentrations, reservoirs in Europe and North America exhibit steeper DO declines compared with other continents. This accelerated decline likely reflects the combined effects of reservoir aging and legacy nutrient loading, which enhance eutrophication and oxygen depletion over time7,14. Additionally, intensified thermal stratification under regional warming reduces vertical mixing and oxygen replenishment, further exacerbating hypoxia risks in temperate reservoirs with strong anthropogenic influence. These findings highlight that even technologically advanced regions are not immune to the escalating challenges of deoxygenation, emphasizing the need for adaptive management strategies tailored to regional hydroclimatic and socio-economic contexts.Our findings of widespread deoxygenation of global reservoirs underscore the need to incorporate oxygen dynamics into reservoir operation and design. Thus, integrating DO monitoring into water resource management frameworks is crucial for safeguarding ecosystem services and ensuring reliable water supplies under conditions of climate warming and increasing anthropogenic pressures13,34,44,45. Oxygen depletion in the hypolimnion not only suppresses microbial CH₄ oxidation, thereby reducing its consumption, but also creates favorable anaerobic conditions that enhance CH₄ production, together leading to elevated methane emissions46,47,48. Given that reservoirs are already recognized as hotspots of methane release, the amplification of emissions under deoxygenated conditions poses significant feedback to the global climate system49. Thus, addressing reservoir deoxygenation is not only essential for local water security but also for mitigating global greenhouse gas budgets.Although current mean DO concentrations in most reservoirs remain above the generally accepted ecological threshold of 5 mg/L5, the observed rates of decline suggest that some regions—particularly in Asia and North America—may approach critical oxygen limits within the coming decades if current trends persist. Given the considerable spatial variability in reservoir morphology, climate sensitivity, and management regimes, we refrain from providing a global first-order estimate of when these thresholds might be crossed, as such extrapolation could introduce substantial uncertainties. Nevertheless, our findings underscore the urgency of developing regional-scale models to predict oxygen depletion timelines and inform early adaptive management strategies for maintaining ecological integrity and water quality under ongoing climate and anthropogenic pressures.In this context, our findings align strongly with multiple Sustainable Development Goals (SDGs). By enabling accurate long-term assessments of reservoir oxygen status, our results contribute to improved water quality management (SDG 6: Clean Water and Sanitation) and protection of aquatic ecosystems (SDG 14: Life Below Water). The observed climate sensitivity of DO emphasizes the urgent need for climate adaptation strategies (SDG 13: Climate Action), while the direct link to greenhouse gas emissions integrates reservoir management into global carbon mitigation efforts. Reliable monitoring also supports sustainable agriculture (SDG 2: Zero Hunger) by improving the understanding and management of irrigation water quality, fosters technological innovation (SDG 9: Industry, Innovation, and Infrastructure) in environmental monitoring, and strengthens international cooperation (SDG 17: Partnerships for the Goals) by providing a global benchmark for reservoir management under climate change. Overall, the long-term perspective presented here highlights reservoir deoxygenation as an emerging global water quality challenge with far-reaching implications. Incorporating oxygen dynamics into reservoir design, operation, and governance frameworks is essential for sustaining aquatic biodiversity, securing safe water supplies, and contributing to global sustainability agendas.ConclusionsThis study develops a robust method for monitoring global reservoir deoxygenation by integrating multi-decadal Landsat observations with machine learning. Our novel dissolved oxygen (DO) retrieval method achieve an R2 is 0.73 and RMSE is 1.23 mg/L in testing set. Using this method, we reveal that 74% of the global major reservoirs exhibited widespread deoxygenation over the past four decades. The widespread decreasing of reservoirs DO drive primarily by climate warming (~ 46%) and human perturbations (~ 31%), including agricultural runoff and land use changes. Our analysis also reveals pronounced spatial heterogeneity in global reservoir oxygenation, with consistently higher DO concentrations in cold northern and polar systems and lower, more variable conditions in tropical and subtropical regions. Temporally, reservoirs undergo a significant global decline in surface DO at a mean rate of − 0.13 mg/L per decade, a pace exceeding that observed in natural lakes, rivers, and oceans. Seasonal fluctuations further amplify this variability, with contrasting hemispheric patterns driven by temperature-dependent solubility and biological oxygen demand. Our satellite-based approach developed here provides a scalable solution for long-term reservoir DO monitoring, underscoring both the climatic sensitivity and ecological vulnerability of reservoir oxygen dynamics. Future advancements in remote sensing, coupled with expanded in-situ validation, will further enhance our ability to track and predict these changes, supporting evidence-based conservation efforts in a warming world.

    Data availability

    Data are available on request. Please contact the corresponding author.
    ReferencesBlaszczak, J. R. Deoxygenation of temperate rivers. Nat. Clim. Chang. 13, 1021–1022 (2023).Article 
    ADS 

    Google Scholar 
    Jane, S. F. et al. Widespread deoxygenation of temperate lakes. Nature 594, 66–70 (2021).Article 
    ADS 
    PubMed 

    Google Scholar 
    Zhi, W., Klingler, C., Liu, J. & Li, L. Widespread deoxygenation in warming rivers. Nat. Clim. Chang. https://doi.org/10.1038/s41558-023-01793-3 (2023).Article 

    Google Scholar 
    Xiao, Q. et al. Lakes shifted from a carbon dioxide source to a sink over past two decades in China. Sci. Bull. 69, 1857–1861 (2024).Article 

    Google Scholar 
    Rose, K. C. et al. Aquatic deoxygenation as a planetary boundary and key regulator of Earth system stability. Nat. Ecol. Evol. 8, 1400–1406 (2024).Article 
    PubMed 

    Google Scholar 
    Kim, Y. H. et al. Application of satellite remote sensing in monitoring dissolved oxygen variabilities: A case study for coastal waters in Korea. Environ. Int. 134, 105301 (2020).Article 
    PubMed 

    Google Scholar 
    Zhang, Y., Shi, K., Woolway, R. I., Wang, X. & Zhang, Y. Climate warming and heatwaves accelerate global lake deoxygenation. Sci. Adv. 11, eadt5369 (2025).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Mooney, B. P., Gårdmark, A., Faithfull, C., Van Dorst, R. M. & Huss, M. Warmer lakes support phytoplankton over fish. Glob. Chang. Biol. 31 (2025).Li, Y., Robinson, S. V. J., Nguyen, L. H. & Liu, J. Satellite prediction of coastal hypoxia in the Northern Gulf of Mexico. Remote Sens. Environ. 284, 113346 (2023).Article 

    Google Scholar 
    Carey, C. C. et al. Anoxia decreases the magnitude of the carbon, nitrogen, and phosphorus sink in freshwaters. Glob. Change Biol. 28, 4861–4881 (2022).Article 
    ADS 

    Google Scholar 
    Lehner, B. et al. High-resolution mapping of the world’s reservoirs and dams for sustainable river‐flow management. Front. Ecol. Environ. 9, 494–502 (2011).Article 

    Google Scholar 
    Li, Y., Zhao, G., Allen, G. H. & Gao, H. Diminishing storage returns of reservoir construction. Nat. Commun. 14, 3203 (2023).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Beck, A. J. & Sañudo-Wilhelmy, S. A. Impact of water temperature and dissolved oxygen on copper cycling in an urban estuary. Environ. Sci. Technol. 41, 6103–6108 (2007).Article 
    ADS 
    PubMed 

    Google Scholar 
    Cai, S. et al. Human-driven nutrient structure shifts as a new driver of coastal deoxygenation: evidence from long-term observations in a Bay with frequent human activities. Mar. Pollut. Bull. 220, 118444 (2025).Article 
    PubMed 

    Google Scholar 
    Breitburg, D. et al. Declining oxygen in the global ocean and coastal waters. Science 359, eaam7240 (2018).Article 
    PubMed 

    Google Scholar 
    Ross, M. R. V. et al. AquaSat: A data set to enable remote sensing of water quality for inland waters. Water Resour. Res. 55, 10012–10025 (2019).Article 
    ADS 

    Google Scholar 
    Virro, H., Amatulli, G., Kmoch, A., Shen, L. & Uuemaa, E. G. R. Q. A. Global river water quality archive. Earth Syst. Sci. Data. 13, 5483–5507 (2021).Article 
    ADS 

    Google Scholar 
    Tu, Z., Zhang, Y., Shi, K., Gong, S. & Zhao, Z. Landsat data reveal lake deoxygenation worldwide. Water Res. 267, 122525 (2024).Article 
    PubMed 

    Google Scholar 
    Beal, M. R. W., Özdoğan, M. & Block, P. J. A Machine learning and remote sensing-based model for algae pigment and dissolved oxygen retrieval on a Small Inland Lake. Water Resour. Res. 60, e2023WR035744 (2024).Article 
    ADS 

    Google Scholar 
    Schmidtko, S., Stramma, L. & Visbeck, M. Decline in global oceanic oxygen content during the past five decades. Nature 542, 335–339 (2017).Article 
    ADS 
    PubMed 

    Google Scholar 
    Oschlies, A., Brandt, P., Stramma, L. & Schmidtko, S. Drivers and mechanisms of ocean deoxygenation. Nat. Geosci. 11, 467–473 (2018).Article 
    ADS 

    Google Scholar 
    Earle, S. A. et al. Ocean deoxygenation: time for action. Science 359, 1475–1476 (2018).Article 
    ADS 
    PubMed 

    Google Scholar 
    Dethier, E. N., Renshaw, C. E. & Magilligan, F. J. Toward improved accuracy of remote sensing approaches for quantifying suspended sediment: Implications for suspended-sediment monitoring. JGR Earth Surf. 125, e2019JF005033 (2020).Article 
    ADS 

    Google Scholar 
    Tian, S. et al. A novel framework for river organic carbon retrieval through satellite data and machine learning. ISPRS J. Photogramm. Remote Sens. 221, 109–123 (2025).Article 
    ADS 

    Google Scholar 
    Jones, J. W. Improved automated detection of Subpixel-Scale Inundation—Revised dynamic surface water extent (DSWE) partial surface water tests. Remote Sens. 11, 374 (2019).Article 
    ADS 

    Google Scholar 
    Pekel, J. F., Cottam, A., Gorelick, N. & Belward, A. S. High-resolution mapping of global surface water and its long-term changes. Nature 540, 418–422 (2016).Article 
    ADS 
    PubMed 

    Google Scholar 
    Vermote, E., Justice, C., Claverie, M. & Franch, B. Preliminary analysis of the performance of the Landsat 8/OLI land surface reflectance product. Remote Sens. Environ. 185, 46–56 (2016).Article 
    ADS 

    Google Scholar 
    Roy, D. P. et al. Characterization of Landsat-7 to Landsat-8 reflective wavelength and normalized difference vegetation index continuity. Remote Sens. Environ. 185, 57–70 (2016).Article 
    ADS 

    Google Scholar 
    Knaeps, E. et al. The SeaSWIR dataset. Earth Syst. Sci. Data. 16. https://doi.org/10.1594/PANGAEA.886287 (2018).Basu, A., Culpepper, J., Blagrave, K. & Sharma, S. Phenological shifts in lake ice cover across the Northern hemisphere: A glimpse into the Past, Present, and the future of lake ice phenology. Water Resour. Res. 60, e2023WR036392 (2024).Article 
    ADS 

    Google Scholar 
    Tian, S. et al. Remote sensing retrieval of inland water quality parameters using Sentinel-2 and multiple machine learning algorithms. Environ. Sci. Pollut Res. 30, 18617–18630 (2022).Article 

    Google Scholar 
    Zhu, X., Guo, H., Huang, J. J., Tian, S. & Zhang, Z. A hybrid decomposition and machine learning model for forecasting Chlorophyll-a and total nitrogen concentration in coastal waters. J. Hydrol. 619, 129207 (2023).Article 

    Google Scholar 
    Chen, T., Guestrin, C. & XGBoost: A Scalable Tree Boosting System. in Proceedings of the 22nd ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, 785–794 (ACM, San Francisco California USA, 2016). https://doi.org/10.1145/2939672.2939785Jansen, J. et al. Climate-driven deoxygenation of Northern lakes. Nat. Clim. Chang. 14, 832–838 (2024).Article 

    Google Scholar 
    Maavara, T. et al. River dam impacts on biogeochemical cycling. Nat. Rev. Earth Environ. 1, 103–116 (2020).Article 
    ADS 

    Google Scholar 
    Xiao, Q. et al. Eutrophic lake Taihu as a significant CO2 source during 2000–2015. Water Res. 170, 115331 (2020).Article 
    PubMed 

    Google Scholar 
    Song, C. et al. A comprehensive Geospatial database of nearly 100 000 reservoirs in China. Earth Syst. Sci. Data. 14, 4017–4034 (2022).Article 
    ADS 

    Google Scholar 
    Tian, S., Guo, H., Huang, J. J., Zhu, X. & Zhang, Z. Comprehensive comparison performances of Landsat-8 atmospheric correction methods for inland and coastal waters. Geocarto Int. 37, 15302–15323 (2022).Article 
    ADS 

    Google Scholar 
    Remer, L. A. et al. Retrieving aerosol characteristics from the PACE Mission, part 2: Multi-Angle and polarimetry. Front Environ. Sci 7 (2019).Limburg, K. E., Breitburg, D., Swaney, D. P. & Jacinto, G. Ocean deoxygenation: A primer. One Earth. 2, 24–29 (2020).Article 
    ADS 

    Google Scholar 
    Kim, H., Franco, A. C. & Sumaila, U. R. A selected review of impacts of ocean deoxygenation on fish and fisheries. Fishes 8, 316 (2023).Article 

    Google Scholar 
    Pitcher, G. C. & Jacinto, G. S. 3.3 Ocean deoxygenation links to harmful algal blooms.Conway, D. et al. Climate and Southern africa’s water–energy–food nexus. Nat. Clim. Change. 5, 837–846 (2015).Article 
    ADS 

    Google Scholar 
    Lee, M. et al. Uneven consequences of global climate mitigation pathways on regional water quality in the 21st century. Nat. Commun. 15, 5464 (2024).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Oschlies, A. A committed fourfold increase in ocean oxygen loss. Nat. Commun. 12, 2307 (2021).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Chen, Z. et al. Extreme warming intensified lacustrine deoxygenation and methane cycling during the Paleocene-Eocene thermal maximum. Commun. Earth Environ. 6, 592 (2025).Article 
    ADS 

    Google Scholar 
    Żygadłowska, O. M. et al. Eutrophication and deoxygenation drive high methane emissions from a brackish coastal system. Environ. Sci. Technol. 58, 10582–10590 (2024).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Xiao, Q. et al. Spatial variations of methane emission in a large shallow eutrophic lake in subtropical climate: methane flux in a shallow eutrophic lake. J. Geophys. Res. Biogeosci. 122, 1597–1614 (2017).Article 

    Google Scholar 
    Maavara, T., Lauerwald, R., Regnier, P. & Van Cappellen, P. Global perturbation of organic carbon cycling by river damming. Nat. Commun. 8, 15347 (2017).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Download referencesFundingThe authors declare that no funds, grants, or other support were received during the preparation of this manuscript.Author informationAuthors and AffiliationsSchool of Earth, Atmosphere and Environment, Monash University, Clayton, VIC, 3800, AustraliaLiangwei LiaoYellow River Water Conservancy Commission Midstream Hydrology and Water Resources Bureau, Jinzhong, 030600, ChinaXinge CaiAuthorsLiangwei LiaoView author publicationsSearch author on:PubMed Google ScholarXinge CaiView author publicationsSearch author on:PubMed Google ScholarContributionsLiangwei Liao: Conceptualization, Methodology, Formal analysis, Validation, Writing – Original Draft, Writing – Review & Editing.Xinge Cai : Software, Investigation, Resources, Data Curation, Visualization.Corresponding authorCorrespondence to
    Liangwei Liao.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 articleLiao, L., Cai, X. Satellites reveal widespread deoxygenation of large global reservoirs from 1984 to 2023.
    Sci Rep 15, 44680 (2025). https://doi.org/10.1038/s41598-025-28265-2Download citationReceived: 19 August 2025Accepted: 10 November 2025Published: 29 December 2025Version of record: 29 December 2025DOI: https://doi.org/10.1038/s41598-025-28265-2Share 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
    KeywordsRemote sensingDissolved oxygenMachine learningReservoirs, global scale More

  • in

    Distinct kinematics and micromorphology for symmetrical rowing and sliding on water in ripple bugs and water striders

    AbstractSemiaquatic bugs evolved two different propulsion mechanisms for their symmetrical rowing: a drag-based propulsion in Veliidae and a surface-tension-based propulsion in Gerridae. However, the comparative leg micromorphology and kinematics underlying these two mechanisms remain underexplored. In this study, we compared leg micro- and nanostructures and kinematics of Rhagovelia distincta (Veliidae), which employs midleg fans as oar-like blades for drag-based thrust, with Gerris latiabdominis (Gerridae), which uses long midlegs for surface-tension-based thrust. R. distincta move their midlegs in short strokes and deployed fans which function as “leaky paddles” with higher anteroposterior rigidity, inferred from seta and setula structure, exploiting drag and potentially lift. Fan protraction into the water appeared to require muscle control, while elastocapillarity may contribute to fan shaping. In contrast, G. latiabdominis exhibited longer strokes with midlegs covered with dense hydrophobic hairs suited for surface-tension-based propulsion. Ventral setae on tarsal section producing surface-tension-based-thrust formed longitudinal rows-and-gaps in both species, with posterior rows particularly robust and nano-grooved in G. latiabdominis. Additionally, both formed ventral beam-like structures from overlapping flat-tipped setae on hindlegs and forelegs which are used for support and sliding. These findings generate new hypotheses for refining models of locomotion on water surface by insects with their micro/nano-morphological diversity.

    Similar content being viewed by others

    Functional morphology of the leg musculature in the marine seal louse: adaptations for high-performance attachment to diving hosts

    Article
    Open access
    23 December 2025

    Observation and analysis of diving beetle movements while swimming

    Article
    Open access
    16 August 2021

    Drag enhancement characteristics and mechanism of the seam plate formed by uniform micro gap arrangement of multiple micro cylinders

    Article
    Open access
    10 April 2025

    IntroductionLocomotion on the water surface presents a unique set of physical challenges that have driven repeated and diverse evolutionary solutions among insects1. The semiaquatic bugs (Gerromorpha) provide an example of adaptive radiation into this novel environment. These insects evolved distinct morphological, behavioral, and anatomical traits that enable movement on the air-water interface2. These adaptations reflect multiple, lineage-specific strategies for solving similar functional problems–offering a model system for studying how alternative solutions involving morphology and behavior evolve under ecological constraints of physical environment–the water surface. However, the details of the co-evolutionary and functional interplay between leg microstructures, stroke kinematics, and thrust mechanics across major independent lineages of this adaptive radiation remain poorly understood. To provide more insights into this question, we compare two major independently evolved solutions for symmetrical rowing in Gerromorpha. Here, we focus on how alternative physical mechanisms–drag-based versus surface-tension-based thrust–are realized through contrasting yet functionally convergent morphologies.Notably, two lineages—Veliidae (a polyphyletic family typically found in fast-flowing streams) and Gerridae (typically inhabiting slow or still waters)—have independently evolved symmetrical backward rowing by midlegs for forward thrust2,3,4. While both rely on midlegs for thrust and fore- and hindlegs for support and sliding, their physical modes of thrust generation differ: Veliidae exploit hydrodynamic drag (and potentially lift), while Gerridae generate thrust primarily through surface tension forces5,6. Despite their independent origins and these contrasting mechanisms, the leg kinematics and microstructural adaptations that support these behaviors have not been systematically compared.Within Veliidae, species such as those in the genus Rhagovelia possess specialized midleg pretarsal structures known as swimming fans, which function as oars2,5,7,8,9. These structures are assumed to be actively controlled via a claw retractor muscle5, though recent observations of isolated fans spreading in water have led to the hypothesis of passive elasto-capillary spreading8,9. The nature of fan manipulation in intact, behaving animals remains unresolved. Similarly, the ventral microstructures on midlegs (involved in thrust) and on fore- and hindlegs (involved in support and sliding) have not been systematically examined in Rhagovelia, though their role in generating surface tension-based forces via water dimples is likely.In contrast, Gerridae species such as Gerris latiabdominis do not possess swimming fans but instead have elongated, hairy midlegs that generate thrust through the creation of asymmetric dimples on the water surface. These midlegs, along with highly hydrophobic ventral surfaces, are critical for surface-tension-based propulsion6. While some aspects of locomotory performance have been compared3, detailed kinematic analyses and high-resolution comparisons of leg micromorphology between Gerridae and Veliidae are lacking. Prior studies have presented images of leg hair arrangements2,10, but a focused comparative analysis across functional leg regions has not been conducted since Andersen (1976).Following the framework set by Andersen (1976, 1982) and Crumière et al. (2016), we compare Rhagovelia distincta (Veliidae) and Gerris latiabdominis (Gerridae), two small-bodied representatives of their respective clades. We examine their midleg microstructures and stroke kinematics to evaluate how different physical thrust mechanisms are supported by contrasting anatomical features. We further investigate the hypothesized passive versus active mechanisms of fan control in R. distincta. Finally, we describe and compare fore- and hindleg microstructures used in support and sliding to identify potential convergences across taxa with different thrust mechanisms. Our results provide a foundation for future work on the functional evolution of water-surface locomotion in Gerromorpha.ResultsOverview of figures and supplementary materialsThe results are presented through figures cited in sequential order for clarity and traceability. Figure 1A1–D6 portrays typical thrust phase in Rhagovelia distincta, followed by corresponding views for Gerris latiabdominis in Fig. 2A1–D6. Figure 3A–S illustrates leg seta types. Figures 4A1–C4 and 5A1–C4 show SEMs of leg microstructures. Figure 6A–E compares contact angles and droplet shapes. Stroke kinematics and interspecific comparisons are shown in Figs. 7A–J and 8A1–D, with measurement summaries in Fig. 9A–H.Supplementary figures follow the same order. Figure S1A–B depicts resting leg posture on water. Figure S2A–H shows leg-surface interaction and dimple formation. Figure S3A–E quantifies fan geometry and angles. Figures S4A–D and S5 capture fan dynamics during strokes. Figures S6A–G and S7A–C detail fan positioning within the tarsal cleft. Figures S8A–E and S9A–E present dissected fan–claw complexes. Figure S10A–S expands seta classification. Figures S11, S12A–H, S13–F, S14A–F describe ultrastructural features of fan setae and claws. Figures S15A–F, S16A–D, and S17A–B highlight ventral hook and spoon setae. Figures S18A–D and S19A–H show fore- and hindleg structures in R. distincta. Figures S20A–F and S21A–C provide analogous views for G. latiabdominis. Figures S22A–D and S23A–E show additional fore- and hindleg adaptations. Figure S24A–E and (a–e) detail droplet behavior and surface wetting (contact angles). Figures S25A–D summarize net force estimates during strokes. Figures S26A–D and S27 present multivariate and theoretical analyses of kinematics and fan leakiness.Behavioral observations: leg use at restR. distincta body is supported on foreleg and hindleg tarsus with minimal contribution from the midleg tarsal tips (Fig. 1A1, B1, C1, D1; Figure S1). Occasionally, a very small portion of the midleg`s fan is protruded from the tarsal tip into the water body through the surface it`s in contact with (Fig. 1B1; Video S3 Part 4). G. latiabdominis body is supported on foreleg tarsus, hindleg tibia and tarsus, and midleg intermediate-distal tibia and tarsus (Fig. 2A1, B1, C1, D1; Video S2), which create dimples without piercing the water surface (dimples cast shadows on the bottom of the container; Fig. 2C1, D1; Video S2 Part 4).Behavioral observations: leg use at locomotionDuring a typical initial thrust phase (Fig. 1), R. distincta moves its midlegs forward–either above the water or lightly contacting the surface–then places the tarsi onto the water (Fig. 1C2), which is associated with fan extension into the water (Fig. 1B2). Alternatively, the midlegs may advance while the tarsal tips remain in contact with the surface and a small portion of the fan protrudes underwater (Fig. 1B6; Figure S2D; Video S3 Part 4). Observations from 110 slow-motion videos (each capturing 1–4 fan opening and closing events) suggest that R. distincta actively controls the timing, extent, and duration of fan protraction and retraction, regardless of midleg position on the water surface (SI Part 2; Figure S4; Video S3). For example, we frequently observed the fan opening and closing without any respective decrease or increase in the gap between the tarsus and the water surface (Fig. 1B1). This disagrees with the recently proposed passive fan actuation hypothesis9, which posits that lowering the tarsus onto the water surface is required to initiate fan unfolding through fan-water elastocapillary interactions (SI Part 2).As the fan rapidly protracts (Fig. 1B2; Figure S2C, D) and the midlegs are pushed backward (Fig. 1C4, C5), the tarsus is simultaneously pressed downward (Fig. 1C3; Figures S2B and S5). This generates growing anteroposterior asymmetrical dimples (Fig. 1B4), visible as shadows that expand from the initial miniscule circles at the tarsal tips to ovals extending distally from the tibiotarsal joints (Fig. 1D4). Contrast to G. latiabdominis (Fig. 2D), R. distincta shows an expanded anterior dimple region (Fig. 1C3), likely caused by water displaced beneath the surface by the fan (Fig. 1B4; Figure S2G, H). Strong strokes can produce surface waves (Fig. 1C4, C5; 17 of 99 strokes; Video S1 Part 1), but all strokes transition into a passive sliding phase, during which the midlegs either disengage from the surface or trail behind with minimal fan protrusion (Fig. 1B6; Video S1). Observed midleg disengagement suggests adhesive forces and surface tension are overcome in this process (Video S3). Throughout, the fore- and hindleg tarsi remain in contact with the surface, providing support during sliding.Duration of fan protraction ranged between 6 and 23 ms (12.3 ± 3.4 ms; n = 69) while the duration of fan retraction ranged between 3 and 15 ms (8.4 ± 2.6 ms; n = 73) (Figures S3 and S4). Shorter durations were often associated with strokes beginning or ending with partially protracted fans, where 2–4 distal setae tips protruded into the water. Fully protracted fans had an average projected area of 0.89 ± 0.04 mm2, radius of 0.85 ± 0.01 mm, and protracted angle of 139.13 ± 3.64° (n = 6; Figure S3A). The longitudinal axis of wetted midleg and protracted fan typically lay in the same plane, slanted at 80.2 ± 6.0° (n = 26; Figure S3E).G. latiabdominis generated thrust without piercing the surface (Fig. 2), using midlegs slightly rotated so that anteroventral gap-row microstructures pressed backward against the water (Figure S2; 27–36 ms into the stroke). Backward movement of the midlegs produced anteroposterior asymmetrical dimples and backward-moving surface waves (Fig. 2D; Figure S2E, F; Video S2). Midleg disengagement occurred when the wetted midleg aligned nearly parallel to the direction of body movement and proceeded gradually from proximal to distal segments. Some instances of disengagement appeared smooth (Video S2), while others suggested adhesive forces and surface tension were overcome during the process (Video S2; Figure S2E, 95 ms). Hindlegs contributed weakly to thrust, as indicated by faintly asymmetrical shadows and small wave bows during the initial stroke phase (Fig. 2D4). Hindleg tarsi and tibiae provided support during sliding, shown by shadows aligned with the body axis (Fig. 2D5, D6). Forelegs also supported the body, particularly near the end of the stroke, as indicated by prominent shadows (Fig. 2D6), except at mid-stroke when support was reduced (Fig. 2D4).Ventral microstructures on legsOverviewThe legs of R. distincta had a less dense hair layer than those of G. latiabdominis. We identified 17 setae (hair) types: five shared by both species, five unique to R. distincta, and seven unique to G. latiabdominis (Fig. 3; Figure S10; Table S3). Our analysis focused on the ventral microstructures of leg segments that interact with the water surface, and respective nanometer sized details specifically in R. distincta (Figs. 4 and 5; Figures S11–S23).Ventral microstructures for thrust generationThe swimming fan of R. distincta (Fig. 4B1, B2; Figures S6–S9 and S11–S14) consists of anterior and posterior claws and a fan made up of 17 (Figures S8, S9) to 21 (Fig. 4B2; Figure S11) setae. Each seta bears setulae along its axis at 8–12 μm intervals, forming a feather-like structure (Fig. 4B3, B4; Figures S12, S13). When protracted, the distance between adjacent setae ranges from ~ 20 to ~ 100 μm, and between setulae from several to ~ 20 μm, with typical setula spacing of 4–10 μm (Figure S13A, B). The fan is anchored at the inner proximal corner of the cleft between the two lobes of tarsomere 3 (Fig. 4B2). The surfaces of the setae, setulae, and claws lack nanogrooves (Figure S13C–F). The setae resemble flat beams or boards (Fig. 4B5; Figure S12C) with near-elliptical (Fig. 4B6; Figure S12D) or slightly triangular (Figure S12E) cross-sections, measuring 2–4 μm × 7–10 μm. The orientation of the narrow edges suggests that they face the water during thrust generation (Fig. 4B5; Figures S12, S13). The setulae are also flat, 1–2 μm wide and 300–700 nm thick, with a hollow center (Figure S12H), and their orientation further suggest that they press against the water with the narrow edges. Transverse sections of fan setae show lamellar outer layers and a central hollow (~ 700 nm in diameter) containing internal rods (~ 400 nm in diameter) (Fig. 4B6; Figure S12). Claw cross-sections are 1.5–2.5 μm thick and consist of external lamellar layers and multiple internal layers with nanofibers (100 nm) and cluster of nanofibers (200–400 nm) running in various directions (Fig. 4B8; Figure S14).Passive elastocapillary expansion of the fan in water (Figures S8 and S9) was observed only when the fan was completely dissected–either with or without the anterior claw–and removed from its natural position in the cleft (SI Part 3B). In contrast, observations of intact fans and claws anchored naturally within the cleft indicate that fan expansion into the water at the onset of use is not passive (Videos S1 and S3; SI Part 3B). Moreover, we were able to induce fan protraction by mechanically pulling the ut tendon connected to the base of the fan-and-claw structure (Figures S8 and S9), supporting the involvement of active muscular control.Along the ventral edge of the posterior lobe (Figures S15A–F and S16A–D), a structure composed of three rows of H2 setae—spaced 6–8 μm apart within each row and separated by two 2.5–5 μm wide gaps—forms a band that presses against the water surface without breaking it during a stroke (Fig. 4B9; Figure S16). This pattern resembles the “gaps and rows” arrangement seen on the tarsus of G. latiabdominis (Fig. 5B). Similarly, the ventral edge of the anterior lobe, which also contacts the water surface during a stroke, is lined with a band of 3–4 rows of H1 setae (Fig. 4B9; Figure S16); however, only lateral views were available, limiting precise row counts. On ventral tarsomere 2, which also interacts with the water surface without breaking it (Figure S5), we observed three rows: a posterior row of H1 setae and two anterior rows of Sp setae, separated by gaps (Fig. 4B12, B13; Figure S17).In G. latiabdominis, a “gaps and rows” arrangement of setae was observed along ventral midleg sections that interact with the water surface (Fig. 5B; Figures S20A–F and S21A–C). This pattern resembles the structure on the ventral edge of the lobe in R. distincta (Fig. 4B9). It is especially prominent on the tarsus, where a main ventral gap separates a row of cuspidate setae (C) from a posterior row of thorn-like T2 setae, with a second T2 row positioned further posteriorly to form a T2–T2 gap (Fig. 5B4–B9). Anterior to the C row, a row of M1 setae creates a C–M1 gap (Figure S20E). The tips of these three setae types bend distally along the leg`s longitudinal axis, particularly the cuspidate setae, whose long, flat distal sections may contact each other. In contrast, setae on the lateral and dorsal leg surfaces are relatively straight (Figure S21B). On tarsomere 2, a second, less regularly arranged anterior M1 row creates an M1–M1 gap (Figure S20C, E). This arrangement is less distinct on the tibia, where T2 setae exhibit intermediate morphology between T2 and M1, and the C setae are replaced by M2 (Fig. 5B3). At the tip of the tarsus, the pattern disappears, replaced by a ventral concentration of grass-blade-like setae (g) (Fig. 5B12).Ventral microstructures for support and slidingIn R. distincta, specialized microstructures involved in support and sliding were found on the ventral tarsi of the forelegs (Fig. 4C), hindlegs (Fig. 4A), and at the distal tips of the midleg tarsus (Fig. 4B10, B11). These structures consist of one or more rows of Sp setae with flattened, bent tips that overlap to form a “beam-like” surface ~ 20–25 μm below the leg cuticle, oriented toward the water. Two Sp rows were observed on the ventral forelegs (Fig. 4C3, C4; Figure S18) and 1 on the hindlegs (Fig. 4A1; Figure S19), with the flattened tips especially prominent near the leg ends (Figs. 4A5, A6; 4C4). Sp rows are flanked—particularly posteriorly—by H1 setae on the forelegs (Fig. 4C; Figure S18) and M2 setae on the hindlegs (Fig. 4A; Figure S19). At the tips of the midlegs, which also provide support, a dense cluster of Sp setae—likely modified H1 types—was observed, with long, flat, overlapping tips (Fig. 4B11; Figure S15F, G).In G. latiabdominis, specialized microstructures involved in support and sliding were found on the ventral forelegs and hindlegs. On the ventral foreleg tarsus, a band of grass-blade setae (g) with overlapping flat, bent distal tips was observed ~ 15–20 μm from the leg cuticle, facing the water surface (Fig. 5C4; Figure S22). An entangled cluster of web setae (W) was also present, especially near the tibiotarsal joint (Fig. 5C2; Figure S22). On the ventral side of the hindleg tibia and tarsus, 2–3 rows of leaf-blade setae (L) were arranged in an orderly manner (Fig. 5A; Figure S23). Their overlapping distal tips formed a “beam-like” surface with nanogrooves, ~ 20 μm above the water surface (Fig. 5A2; Figure S23), and were accompanied by a posterior row of large thorn setae (T3) (Fig. 5A6, A8; Figure S23A, E). The ventral sides of joints were covered with bundles of L and leaf-like l setae (Fig. 5A1, A6), with l setae also present at the tarsal tips (Fig. 5A9).Contact angle on midleg sections used in thrust generationMidleg tarsal surfaces in G. latiabdominis were generally more hydrophobic than those in R. distincta (Fig. 6). In R. distincta, water droplets rapidly lost their spherical shape after contacting the hair layer, spreading 10–40% and showing decreased shape indices. Contact angles on the dorsal and ventral midleg surfaces progressively decreased from approximately 130.3° and 75.5°, respectively, to as low as 21.9° as droplets collapsed, indicating relatively high surface wettability (Fig. 6A, B; Figure S24). This effect was especially pronounced on the ventral side, where the swimming fan and associated microstructures interact with water during locomotion (Fig. 6B, C; Figure S24). Contact angle measurements revealed that the ventral side of R. distincta`s midleg—particularly the region housing the swimming fan—exhibits clear hydrophilic properties, in contrast to the more hydrophobic surfaces observed in G. latiabdominis. This elevated wettability likely facilitates water surface penetration during fan deployment, reducing resistance and promoting stable submersion of the fan. Such localized wetting, combined with the fan`s structural stiffness, inferred from setae and setulae microstructure, may enhance the effectiveness of drag-based thrust generation. In contrast, droplets on the dorsal and ventral midleg surfaces of G. latiabdominis retained their spherical shapes with minimal spreading (shape index ≈ 1.5), and contact angles remained high throughout dissipation, ranging from 132.1° to 109.3°, consistent with strong hydrophobicity (Fig. 6D, E; Figure S24; SI Part 4). These uniformly hydrophobic surfaces align with G. latiabdominis` reliance on surface tension, supporting the leg`s ability to retain air layers and resist surface penetration during sliding or thrust.Kinematics profiles during a strokeIn both species (Fig. 7), the midleg femur angle increased gradually during the stroke, but more steeply in R. distincta than in G. latiabdominis. The initial angle was more acute in R. distincta (~ 20°) compared to G. latiabdominis (~ 60°), with both reaching a final angle of ~ 120° (Fig. 7A, B). Peak femur angular velocity was higher in R. distincta and occurred mid-stride at a femur angle of 85°, whereas in G. latiabdominis it peaked at 100° (Fig. 7A, B). In G. latiabdominis, the femur–tibia and tibia–tarsus angles remained relatively small throughout the stroke. In contrast, R. distincta showed more pronounced changes: the femur–tibia angle increased from 30° to 60° (Fig. 7C), and the tibia–tarsus angle ranged from 8° to 16° (Fig. 7D). These patterns suggest that R. distincta engages in coordinated rotations at the coxa–femur, femur–tibia, and—though to a lesser extent—tibia–tarsus joints, while in G. latiabdominis, motion is largely concentrated at the coxa–femur joint.In both species, leg velocity during slower (longer) strokes remained below the theoretical critical velocity (~ 0.23 m/s; Fig. 7E), which marks a transition threshold above which thrust generating legs of semiaquatic organisms generate capillary-gravity waves to achieve higher propulsion force11,12,13. In faster (shorter) strokes, leg velocity exceeded this threshold within ~ 5 ms and reached higher peak values in G. latiabdominis than in R. distincta (Fig. 7E). In R. distincta, the leg velocity vector was briefly aligned with the body movement axis only during mid-stroke, when the femur was approximately perpendicular to the trajectory (Figs. 7G and 9G). For the remaining stroke, the fan–acting as an oar blade–moved in a direction misaligned with the body axis and not perpendicular to its movement direction. In contrast, in G. latiabdominis, the leg velocity vector remained nearly parallel to the body movement axis (Figs. 7F and 9E) and nearly perpendicular to the wetted leg section (Figs. 7H and 9G) for a much larger portion of the stroke (rectangles on x-axis). In both species, the highest net force per stroke, indicated by peak body acceleration (Fig. 7J), coincided with intervals of high “effectiveness” indices (Fig. 7F–H; SI Part 5).Kinematics comparisons of a strokeA small subset of G. latiabdominis strokes showed notably higher average leg linear velocities, body velocities, and peak accelerations (Fig. 8). However, when all strokes were analyzed together, there were no significant differences between the two species in average midleg velocity (Fig. 8B2; Table S4), final body velocity (Fig. 8A1; Table S4), or maximal body acceleration (Fig. 8A2; Table S4). Strokes by R. distincta were shorter in duration (Fig. 8C3; Table S4) and involved faster angular femur movements (Fig. 8B1; Table S4). The legs traveled a shorter distance across the water surface during each stroke compared to G. latiabdominis (Fig. 8C1, C2). This difference is notable because R. distincta has shorter legs (Tables S1 and S2), and additional angular movements were observed at the tibia and tarsus (Fig. 7C, D). Since the kinematic variables are intercorrelated (Figure S26D), estimates of their individual effects on body speed (Figure S26A–C; Tables S5, S6 and S7) are not independent. To address this, we performed a principal component analysis and extracted two components (Table 1). Strokes by R. distincta were characterized by shorter duration and higher femur angular velocity (higher RC2 values), along with shorter distance and slower leg speed on the water surface (lower RC1 values) (Fig. 8E), distinguishing them from G. latiabdominis.Estimated force output for thrust generationAlthough we could not directly measure resistance forces during sliding and thus cannot determine total thrust, the body acceleration profiles during stokes allowed us to estimate the net horizontal thrust vector and the peak net force generated per stroke (Figure S25). Given that approximately 85–95% of total thrust typically translates into forward body momentum in water striders12, our net force estimates likely underestimate the true thrust force by 5–15%.In R. distincta, the maximum horizontal net force per stroke averaged 152 µN with an absolute maximum of 324 µN (Figure S25C). This force was generated through the symmetrical action of two pretarsal swimming fans, each with a projected area of 0.89 ± 0.04 mm2 (1.78 mm² total; Figure S3), along with contribution from wetted tarsi measuring 1.80 ± 0.06 mm (Table S1) in length. Assuming the primary thrust originates from the fans, this translates to an average force of 85 µN/mm² of fan surface area per stroke with an absolute maximum of 183 µN/mm².In G. latiabdominis, the maximum horizontal net force per stroke average 360 µN with an absolute maximum of 745 µN (Figure S25D). This force was generated using wetted midlegs that depress the water surface to create dimples, without piercing it. Based on average wetted leg length of 7.76 ± 0.11 mm (Table S2), this corresponds to an average of 24 µN/mm of leg length with an absolute maximum of 48 µN/mm. To compare per-unit-area force outputs, we approximated the wetted leg segment as a cylindrical surface. With a midleg diameter ranging from 80 to 110 μm (Figures S20 and S21), we used an average radius of 0.05 mm. Assuming half the lateral surface of a 1 mm-long cylinder interacts with the water, the effective thrust-generating area is ~ 0.157 mm2. Based on this, G. latiabdominis generated an average of 148 µN/mm² of interacting leg surface with an absolute maximum of 305 µN/mm².These estimates reflect functional differences in thrust-generation strategies: drag-based propulsion in R. distincta through fan employment and surface-tension-based propulsion in G. latiabdominis through longitudinal row-and-gaps ventral setal structures engagement. Importantly, these values offer a quantitative baseline for future mechanical or computational modeling efforts aimed at linking leg microstructure to propulsion performance.Table 1 Principal component analysis of behavioral variables. PCA was conducted on seven behavioral variables from R. distincta (n = 21) and G. latiabdominis (n = 12). The table shows eigenvalues, percentage of variance explained, and loadings for the first two rotated components (RC1 and RC2), based on the fa.parallel and principal functions from the psych R package. Loadings with absolute values greater than 0.75 are shown in bold. Related results are illustrated in Fig. 8D.Full size tableFig. 1Rhagovelia distincta during the thrust phase. (A1–A6) Side view above the water surface, showing interactions between the leg and water surface. (B1–B6) Side view below the surface, highlighting the motion of the swimming fan during the thrust phase of a stroke; (B1) Examples of fan opening and closing that are not associated with changes in gap size between water surface and distal surface illustrating how fan protracts and retracts without changes of tarsal position relative to water surface (SI Part 2; Figures S4 and S5). (C1–C6) Top view capturing body and leg positions throughout the thrust phase. (D1–D6) Bottom view from beneath the container, showing shadows cast by the body and water-surface dimples. Abbreviations: Fe – femur; Tb – tibia; T1–T3 – tarsomeres 1–3; WL – wetted midleg length; Wb – wave bow. Panels B4 and C3: blue and green annotations indicate interpreted differences in dimple shape based on comparisons with G. latiabdominis (see Fig. 2 and Figure S2).Full size imageFig. 2Gerris latiabdominis during the thrust phase. (A1–A6) Side view above the water surface, showing midleg motion during thrust. (B1–B6) Side view above the water, focused on midleg interaction with the water surface. (C1–C6) Top view capturing body and leg positions throughout the thrust phase. (D1–D6) Bottom view from beneath the container, showing shadows cast by the body and water-surface dimples. Abbreviations: Tb – tibia; T1–T2 – tarsomeres 1 and 2; WL – wetted midleg length; Wb – wave bow.Full size imageFig. 3Schematic drawings of different types of setae found on leg sections of Rhagovelia distincta and Gerris latiabdominis that interact with the water surface. (A) Microsetae, m. (B) Macrosetae 2, M2. (C) Grooming comb, G. (D) Cuspidate setae, C. (E) Stumped setae, S. (F) Macrosetae 2, M2. (G) Hook setae 1, H1. (H) Hook setae 2, H2. (I) Spoon setae, Sp. (J) Obtuse setae, O. (K–L) Macrosetae 1, M1. (M) Thorn setae 1, T1. (N) Thorn setae 2, T2. (O) Thorn setae 3, T3. (P) Web setae, W. (Q) Leaf-blade setae, L. (R) Leaf-like setae, l. (S) Grass-blade setae, g. Alphabetical labels correspond to SEM photos in Figure S10 and morphological data in Table S3. Setae found in both species are labeled in black; those found only in R. distincta are in blue; and those found only in G. latiabdominis are in green. Note that scales vary between panels and are specified for each seta type. Descriptions of all seta types are provided in Supplementary Information Part 3 C.Full size imageFig. 4Scanning electron microscopy summary of leg microstructures in Rhagovelia distincta. (A) Hindleg: (A1) a row of spoon setae, Sp, on proximal ventral tarsomere 2; (A2) anterio-ventral view of the tarsal joint (between tarsomeres 1 and 2), showing overlapping Sp setae with flattened tips, flanked by macrosetae 2, M2; (A3–A5) anterio-ventral views from proximal tarsomere 2 to the tarsal tip, showing a progressively beam-like structure formed by overlapping Sp setae on the ventral side; (A6) a row of spoon setae, Sp, on distal tarsomere 2; (A7) posterior-lateral view of distal tarsomere 2 with a row of spoon setae, Sp, along the water-interacting ventral side (yellow shading in A2–A7). (B) Midleg: (B1) schematic of the pretarsal swimming fan used for hydrodynamics-based thrust; (B2) anterio-ventral view of tarsomere 3 showing the fan and anterior claw extending from the cleft between two lobes (posterior claw not visible); (B3) protracted fan in water showing hierarchical structure of setae and setulae; (B4) anterior view of folded fan, highlighting relative thickness of setae and setulae; (B5) “board-like” cross-sectional shape of fan; (B6) cross-section of a fan seta showing internal layers, hollow core with pillars, and outer layers; (B7) surface of the anterior claw extruding from the cleft surrounded by hook setae (anterior lobe`s H1 setae visible); (B8) cross-section of the claw; (B9) Ventral edges of the cleft with rows of H1 (anterior lobe) and H2 (posterior lobe) setae; (B10) anterior view of the distal portion of tarsomere 3; (B11) close-up of the anterior lobe tip showing long, flattened modified H1 setae resembling spoon setae, Sp, and internal cleft wall lined with H2; tip of anterior claw also visible (posterior lobe removed); (B12) lateral view of tarsomere 2; (B13) ventral view of tarsomere 2 showing orderly rows of Sp and H1 setae. (C) Foreleg: (C1) posterior view showing sparse cuspidate setae, C, dorsal macrosetae 2, M2, and ventral hook setae, H1; (C2) anterio-ventral view showing microsetae, m, ventral hook setae, H1, and a row of spoon setae, Sp along the ventral water-interacting surface; (C3) close-up of spoon setae, Sp, near the claw base; (C4) array of spoon setae, Sp, at the tarsal tip. Color-shaded regions in SEM panels denote ventral (water-interacting) leg surfaces. Additional SEMs are provided for hindlegs in Figure S19, midlegs in Figures S11–S17, and forelegs in Figure S19.Full size imageFig. 5Scanning electron microscopy summary of leg microstructures in Gerris latiabdominis. (A) Hindleg: (A1) ventral view of femorotibial joint covered with leaf-blade setae, L; (A2) ventral proximal tibia showing overlapping distal sections of L setae forming a “beam-like” structure believed to support the insect on water; (A3) same as A2, but from a different preparation; L setae appear more randomly bent due to cleaning and drying procedures. A row of large thorn setae 3, T3, runs posterior to the L setae, with macrosetae 2, M2, present on the leg`s posterior side; (A4) lateral view of distal tibia with dense microsetae, m, macrosetae 1, M1, thorn setae 1, T1, and a ventral row of thorn setae 3, T3, on the water-interacting surface; (A5) ventral view of tarsomere 1 with a longitudinal row of leaf-blade setae, L, and an adjacent row of thorn setae 3, T3; (A6) ventral view of tibiotarsal joint with arrays of leaf-like setae, l, and thorn setae 3, T3, extending from distal tibia; (A7) ventral view of tarsomere 2 with a continuing row of leaf-blade setae, L, and an adjacent row of thorn setae 3, T3. (A8) close-up of ventral overlapping L setae forming a flat “beam-like” surface with nano-grooves running longitudinally; (A9) ventral view of tarsal tip with an array of leaf-like setae, l. (B) Midleg: (B1) overview of the “gaps and rows” arrangement on ventral midleg segments involved in thrust generation (purple shading). Top panel: full tibia and tarsus; middle panel: tarsus with one main longitudinal gap and two less distinct gaps; bottom panel: clearer visualization of the three gaps, each flanked by linear setal rows. (B2) ventral view of intermediate tibia with macrosetae 2, M2, flanked by rows of macrosetae 1, M1, separated by noticeable gaps; (B3) distal tibia with continuing macrosetae 2, M2, and adjacent rows of thorn setae 2, T2, which resemble M1; (B4–B6) anteroventral view of tarsomere 1, with a row of cuspidate setae, C, adjacent rows of thorn setae 2, T2, and a gap in between; (B7–B9) ventral view of tarsomere 2 with a continuing row of cuspidate setae, C, and two posterior rows of thorn setae 2, T2; visible are the main gap between C and T2, and a narrow gap between the two T2 rows; (B10) anterior view of tarsomere 2 (B11) clear view of the main gap between C and T2 rows; (B12) tarsal tip with an array of grass-blade setae, g. (C) Foreleg: (C1) grooming comb, G, and stumped setae, S, on dorsal tibiotarsal joint; (C2) anterior-lateral view of proximal tarsus with stumped setae, S, on lateral side, and overlapping grass-blade, g, and web setae, W, on the ventral water-interacting surface; (C3) posterior-lateral view of distal tarsus with grass-blade, g, and web setae, W; (C4) close-up of overlapping distal tip of grass-blade, g. Color-shaded regions in SEM panels denote ventral (water-interacting) leg surfaces. Additional SEMs are provided for hindlegs in Figure S23, midlegs in Figures S20 and S21, and forelegs in Figure S22.Full size imageFig. 6Contact angle and droplet shape on distal tarsus of Rhagovelia distincta and Gerris latiabdominis. Top panels are example images of water droplets on different leg surfaces. Middle panels contain corresponding SEM images showing dorsal hair layers (A, D), ventral hair layers (B, E), and the surface of R. distincta`s claw (C). Bottom panels are time-course changes (relative to droplet disappearance, set as 100%) in droplet shape index (height/width, yellow squares) and contact angle (degrees) measured on the proximal (orange diamonds) and distal (purple circles) sides of the droplet. (A) Contact angle on R. distincta`s dorsal midleg tarsus (tarsomere 3). (B) Contact angle on R. distincta`s ventral midleg tarsus (tarsomere 3). (C) Contact angle on R. distincta`s claw. (D) Contact angle on G. latiabdominis`s dorsal midleg tarsus (tarsomere 2). (E) Contact angle on G. latiabdominis`s ventral midleg tarsus (tarsomere). For additional details, see Figure S24 and commentary in SI Part 4, and Video S4.Full size imageFig. 7Kinematic profiles of Rhagovelia distincta and Gerris latiabdominis during a stroke. (A) Midleg femur angle (degrees ± SE): angle between femur and body axis; (B) Midleg angular velocity (degrees/s ± SE), derived from A. (C) Femur-tibia angle (degrees ± SE). (D) Tibia-tarsus angle (degrees ± SE). (E) Leg velocity (mm/s ± SE) along the direction of the wetted midleg trajectory on the water surface. (F) “Effectiveness” of leg velocity vector`s direction. (G) “Effectiveness” of leg length`s use. (H) “Effectiveness” of wetted leg orientation. (I) Body velocity (mm/s ± SE) along the body movement axis. (J) Body acceleration (mm/s2 ± SE), derived from I with spline curve fitted to the averages. Insets in (A–D) illustrate how each angle was measured: dark blue indicates fast strokes of R. distincta, light blue indicates slow strokes of R. distincta, dark green indicates fast strokes of G. latiabdominis, and light green indicates slow strokes of G. latiabdominis. Period of high “effectiveness” in (F and H) and (G) illustrate range of ± 0.1 and ± 10°, respectively, from each maximal “effectiveness” observed. Gray-shaded regions indicate time intervals where data from all individuals were included in the average; standard error is shown only when n > 4. Sample sizes: 21 strokes from six individuals of R. distincta and 12 strokes from six individuals of G. latiabdominis.Full size imageFig. 8Comparative kinematics between Rhagovelia distincta and Gerris latiabdominis during a single thrust stroke. Each data point represents one of 21 (R. distincta) and 12 (G. latiabdominis) stroke observations from six individuals per species. (A) Body movement variables: (A1) final body velocity (mm/s); (A2) maximal body acceleration (mm/s2). (B) Midleg movement velocity variables: (B1) average femur angular velocity (degrees/s); (B2) average leg velocity on the water surface (mm/s). (C) Midleg movement distance and time variables: (C1) stroke amplitude of the midleg (mm); (C2) distance traveled by the wetted midleg (mm); (C3) stroke duration (s). (D) Principal Component Analysis (PCA): scatterplot of strokes based on two principal components (Table 1) extracted from variables in panels A–C. Ellipses represent 95% confidence intervals: R. distincta in blue, G. latiabdominis in green. RC1 corresponds primarily to distance and speed; RC2 to stroke duration and angular velocity. Shaded bands on x and y axes illustrate the ranges of observed values for each species.Full size imageDiscussionObservations of live R. distincta suggest that thrust during a stroke results from a combination of two forces: hydrodynamic forces generated by the oar-like motion of the fan, as proposed in previous studies5,7,9, and additional capillary forces arising from an anteroposteriorly asymmetrical dimple beneath the tarsus, consistent with surface-tension-based mechanisms described in Gerridae6. In contrast, behavioral evidence from G. latiabdominis aligns with the surface-tension-based thrust mechanism characterized in detail for Aquarius paludum6. The observed differences between these species appear closely tied to variations in leg microstructures, respective internal nano-structural properties, and motion kinematics that reflect their distinct ecological contexts and thrust-generation strategies.In R. distincta, several microstructural features suggest specialization for hydrodynamic thrust. The hydrophilic properties of the claw likely facilitate surface penetration during stroke initiation, as the fan and claw extend downward from the cleft`s internal compartment. Similar surface textures between the claw and fan setae indicate that the fan may also be hydrophilic, enhancing its ability to submerge through the water surface. The fan`s setae and setulae are oriented to press against the water with their narrow edges, a configuration that minimizes deformation under hydrodynamic forces, as predicted by beam theory14 (SI Part D).The internal architecture of the fan setae and setulae appears well-adapted for their role as underwater oars in thrust generation. Each seta comprises a hollow core reinforced with columnar nanofibers, while the associated claw features a lamellar structure—both resembling engineered designs such as sandwich and lamellar composites15,16, which are known to enhance stiffness and fatigue resistance. Polymer nanofibers are widely recognized for their high strength due to a high surface-area-to-volume ratio, and thinner fibers are particularly associated with greater flexibility and mechanical resilience17,18. In R. distincta, the small diameter of nanofibers within the internal walls of these flat and hollow beams likely contributes to their flexibility. Combined with the hydrophilic properties of the surface, this structure may underlie the elastocapillary behavior responsible for the fan`s characteristic “fan-like” shape when immersed in water—an effect observed both in our experiments (SI Part 3B; Figures S8 and S9) and in Ortega-Jimenez et al. (2025). Furthermore, the multidirectional growth pattern of nanofibers likely reinforces the setae across multiple planes, enhancing their ability to resist water forces. As with synthetic hollow nanofiber systems19, the central hollow core may facilitate stress distribution, controlled deformation under load, and reduced material weight—all contributing to the fan`s structural integrity and functional performance resulting in net thrust outputs of ~ 300 µN per stroke and 180 µN/mm2 of fan surface area.Additionally, the H1 and H2 setae rows positioned at the cleft entrance likely act as a barrier against water intrusion while assisting fan deployment through elastocapillary interactions. Based on the observed stroke speed, setula spacing, and thickness, we estimate the fans to operate at Reynolds numbers (dimensionless quantity expressing the ratio of inertial to viscous forces in a fluid) ranging from ~ 0.03 to 0.20 (Table S8) and fan`s leakiness (degree to which fluid flows through, rather than around, a porous or bristled structure, depending on geometry and flow regime) to range from ~ 0.3 to 0.6 (Figure S27), suggesting it behaves as a “leaky paddle” rather than a solid blade. This is roughly like bristled appendages in copepods and barnacle larvae20,21. Comparable fan-like structures with potential “leaky paddle” functionality are found in other Veliidae genera such as Tetraripis and Trochopus. In contrast, “Veliidae” species that depend primarily on surface-tension-based thrust, like Velia sp., tend to exhibit more developed ventral “gaps and rows” arrangements and less-developed fan structures5, supporting the hypothesis of divergent functional adaptations.In G. latiabdominis, a different set of microstructural adaptations supports thrust generation primarily through surface tension. The species exhibits denser and more hydrophobic setae on its midlegs than R. distincta, likely reflecting its reliance on surface-tension-based propulsion. Deep, asymmetrical dimples beneath the wetted portion of the leg, along with prominent bow waves, contribute to increased thrust. Interestingly, even the lower hydrophobicity observed in R. distincta can still support thrust via surface tension through dimple formation, though to a lesser degree.Across both species, leg surfaces involved in surface-tension-based thrust display linear arrangements of distally bending setae forming “gaps and rows.” These are absent from other leg surfaces and may serve specialized functions, potentially related to air retention during dimple formation, which prevents surface penetration and facilitates thrust. Theoretical work22,23 suggests such arrangements can trap air and maintain smooth water contact, providing theoretical support for our hypothesis that these structures appear to function like pressurized air pockets that resist water surface breakage under thrust loads. Additionally, the smoother longitudinal gaps may reduce adhesion during stroke recovery. This arrangement is more pronounced in G. latiabdominis, where the setae also exhibit nanogrooves known to enhance hydrophobicity24,25. The posterior concentration of thicker T2 and T3 setae in mid- and hindlegs may be a specific adaptation to withstand the higher pressure during backward strokes, which is exemplified by higher thrust output per leg surface area in G. latiabdominis than R. distincta. These structural reinforcements are aligned with previous findings on jumping and propulsion in surface-dwelling insects26,27.Kinematic data further underscores how thrust generation mechanisms are integrated with species-specific leg movement strategies. In R. distincta, hydrodynamic thrust is supported by a midleg stroke that begins from a more acute femur angle and involves backward rotations across multiple leg joints. The resulting fan movement vectors deviate from the body movement axis similar to patterns characteristic to Xenopus frogs, known to incorporate lift28. This suggests potential contribution of lift-like forces to thrust, akin to paddling strategies in human kayaking29,30 and animal locomotion28,31. The observed combination of Rhagovelia`s leg movement pattern and strokes of shorter duration would be inefficient for surface-tension-based propulsion on stagnant waters but is well-suited to fast-flowing environments where high-frequency of short strokes with minimal surface contact time is advantageous. The ability of R. distincta to actively control fan protraction and retraction through muscle action5 (SI Parts 2 and 3B) could allow for greater flexibility during maneuvering and stroke timing, compared to the hypothetical passive fan deployment proposed in recent studies of another Rhagovelia species8,9, but not supported by our observations.Conversely, G. latiabdominis initiates midleg movement at a less acute femur angle, with primary rotation occurring at the coxa–femur joint, and maintains a nearly straight femur–tibia segment. This configuration enhances wetted surface area, allowing for more efficient surface-tension-based thrust. The leg stroke is longer in duration and follows a nearly parallel trajectory to the body axis, facilitating the formation of asymmetrical dimples critical to curvature force production. The backward movement of the leg, nearly perpendicular to its axis, promotes directional asymmetry in the dimple, optimizing forward propulsion typical for this species.Forelegs and hindlegs support the insect body during thrust and sliding, and both species share a distinct ventral setal arrangement that likely contributes to standing and sliding on the water surface. Flattened, overlapping setae form a beam-like structure along the underside of wetted legs, similar to those described in Gerris and Aquarius5,10. This configuration minimizes surface penetration and drag, offering support without water breakage. The beam`s alignment with body motion also likely facilitates sliding and steering. In the heavier G. latiabdominis, these beams feature hydrophobic nanogrooves that enhance their supporting function. In contrast, the lighter R. distincta lacks such grooves, suggesting lower support demands. Moreover, the spoon-like setae at the tips of R. distincta`s midlegs may help resist displacement by currents, offering a stabilizing advantage in fast-flowing habitats.ConclusionsOur results demonstrate that Rhagovelia distincta and Gerris latiabdominis, despite both having independently evolved symmetrical rowing, employ fundamentally different thrust-generation strategies: R. distincta relies primarily on hydrodynamic drag (and possibly lift) via actively controlled pretarsal fans functioning as “leaky paddles,” while G. latiabdominis generates thrust through surface tension. Detailed morphological and kinematic analyses revealed that fan protraction and retraction in R. distincta involve muscle action, whereas elastocapillarity contributes only to fan conformation once submerged9. The fan`s nanostructured architecture supports this function through its stiffness, flexibility, and surface wettability. We also show that both species exhibit morphological differences linked to their respective thrust strategies and similarities in ventral microstructures—such as longitudinal setal rows-and-gaps and beam-like setal structures—linked to the mechanisms of support and sliding on water surface. These findings challenge simplified models of leg micromorphology in surface-dwelling insects and suggest that precise microstructural arrangements are critical to locomotor performance. By providing detailed, testable hypotheses on structure-function relationships, our study lays a foundation for future mechanical modeling and for comparative evolutionary analyses connecting microstructure, movement, and ecological adaptation in semiaquatic bugs.MethodsField sites and study speciesIn January and February of 2020 and 2023, specimen collections and detailed observations of Rhagovelia distincta (body weight: 4–14 mg; Fig. 9H; Table S1) were made at the Southwestern Research Station, Arizona, USA (SWRS; 31°53′3′′N, 109°12′21′′W). In August and September of 2020, specimen collections and detailed observations of Gerris latiabdominis (16–19 mg; Fig. 9H; Table S2) were made at Gwanak Mountain, Korea (37°26′42′′N, 126°57′51′′E) and Seoul National University, Korea (37°28′57′′N, 126°96′04′′E), respectively. Each individual was weighed (GEM20 High Precision Digital Milligram Jewelry Scale, Smart Weigh, 0.001 g).Videographic and photographic observationsWe filmed four types of high speed and standard videos with Sony RX10-III at 959.04 frames per second (fps) and with Chronos 2.1-HD at 1000–4000 fps of individuals in acrylic containers (18 × 18 cm, filled with water). These standardized still-water conditions were maintained to enable direct, quantitative comparison of kinematic and microstructural variables between species. The strength of this approach is that it removed environmental variability and allowed the isolation of species-specific locomotor traits shaped by their respective ecological conditions.

    Type 1: directly from above (85 and 62 movies collected from six and seven individuals of R. distincta and G. latiabdominis, respectively).

    Type 2: from the side of various angles (below surface, surface level, and above surface) (249 and 87 movies of R. distincta and G. latiabdominis, respectively).

    Type 3: directly from below with light source positioned directly above the container (20 and 60 movies from two and five individuals of R. distincta and G. latiabdominis, respectively) to visualize the shadows on the bottom of the container; shadows correspond to dimples under legs on the water surface.

    Two variables were extracted from the video types 2 and 3:

    Wetted midleg length (mm): maximal leg section in contact with the water surface in the middle of fast thrust strokes when midleg angle to body axis approximates 90°. The wetted midleg consisted of tarsus in R. distincta (6 individuals) and proximal tibia to tarsal tip in G. latiabdominis (6 individuals).

    Projected swimming fan area (mm2): fan surface area was estimated from six still frames taken from six type 2 clips of R. distincta, where the midleg tarsus was perpendicular to both the body axis and camera lens axis (Figure S3). Projected area was calculated assuming the fan forms a circular sector with radius equal to the average of six measured setal lengths. As the fan surface is tilted ~ 80° to the lens (rather than 90°), projections slightly underestimate true area. However, since tilt angles for individual frames were unavailable and the bias is minor, we used the projected area as our estimate.

    Microscopic observations of leg microstructuresUsing optical microscopy, we observed the morphology and behavior of the fan in specimens of R. distincta. Using Scanning Electron Microscopy (SEM), we visualized the hair structures on leg sections that interact with water (SI Part 3).Contact angle measurements (SI part 4)Contact angle (in degrees; °), height and width (mm), and respective shape index (height to width ratio) of small droplets (0.096 ± 0.032 mm) on the surface of ventral and dorsal microstructures of tarsomere 3 and tarsomere 2 of R. distincta and G. latiabdominis, respectively, and on the tarsal claw of R. distincta, were measured (with ImageJ 1.53t) in frames of three high-speed videos per condition (2000 fps; Chronos 2.1-HD Camera, Kron Technologies). The specimens, sprayed with water, were mounted on a micromanipulator (MM-3, Narishige, Japan) parallel to the camera (Video S4).Kinematic profiles of a strokeDetailed kinematic analyses of symmetric strokes by R. distincta and G. latiabdominis were restricted to data extracted with Tracker (https://physlets.org/tracker/) from selected videos: 21 and 12 strokes from six and six individuals for R. distincta and G. latiabdominis, respectively (Supplemental data 1; Supplemental data 2). All strokes were chosen based on strict inclusion criteria—specifically, individuals had to remain stationary before initiating a straight, uninterrupted thrust—to ensure consistent, high-quality data suitable for accurate 2D digitization and meaningful interspecific comparisons. Cartesian (x, y) coordinates of 10 and 9 points on the insect body for R. distincta and G. latiabdominis, respectively (midleg tips were not digitized in G. latiabdominis due to resolution issues) were digitized and subsequently smooth-splined using the “stats” package32,33 (df = 5 and smoothing parameter = 0.5).To compare the two species, we focused on five aspects (Fig. 9) of leg movements during a stroke and extracted the following kinematic variables for each frame, or two consecutive frames, through the thrust phase of a stroke:

    Midleg femur angle (degrees): the angle between the body axis and the femur at the coxae (Fig. 9B) was calculated at each frame. The coxal joint is where the major leg angular movement is performed in both species.

    Femur-tibia angle (degrees): the angle between the femur and the tibia at the femorotibial joint was calculated at each frame (Fig. 9C).

    Tibia-tarsus angle (degrees): the angle between the tibia and the tarsus at the tibiotarsal joint was calculated at each frame (Fig. 9D).

    Midleg angular velocity (degrees/s): calculated by dividing the between-frame difference in midleg femur angles by the latency between the two consecutive frames (i.e., 1/fps).

    Leg velocity (mm/s) (U): the linear velocity in horizontal plane of the midpoint of the wetted midleg length. It is calculated for each pair of two consecutive frames via dividing the distance traveled between by the latency by the two consecutive frames.

    Three proxies of “effectiveness” of midleg application for thrust generation during a stroke were calculated (Fig. 9) using basic trigonometry and vector algebra:

    “Effectiveness” of leg velocity vector`s direction (Fig. 9E) (proportion; range 0–1): the proportion of the leg velocity vector (and of R. distincta`s fan protracted under the leg; green vector) employed along the direction parallel to the body movement (blue or violet vectors). Values closer to ‘1’ indicate “more effective” employment of legs on water surface because the backward leg velocity vector is near-parallel to the body axis line ((:{upalpha:}=0^circ:)) resulting in anteroposterior asymmetry of the dimple crucial for curvature force (i.e., surface tension) contribution to thrust6. Positive values indicate backward velocity vector (blue) that contributes to forward thrust, while negative values indicate forward vector (violet; when legs are dragged along body movement).

    “Effectiveness” of leg length`s use (Fig. 9F) (proportion; range 0–1): evaluation of the relative length of wetted leg projection (blue) on the line perpendicular to the body movement axis (relative to the actual wetted leg length marked green). In G. latiabdominis, it may be approximately viewed as the effective proportion of the total wetted midleg length (blue) that pushes the surface dimple directly backwards along the leg velocity vector parallel to the body movement direction (blue vector in Fig. 9E). Values closer to ‘1’ indicate “more effective” employment of the midleg length pushing the dimple backwards; they also indicate that the fan surface in R. distincta is approximately perpendicular to the body axis ((:{upbeta:}approx:90^circ:)), under the assumption that tarsus on the water surface lies approximately within the near-vertical plane with the surface of the R. distincta`s fan under water.

    “Effectiveness” of wetted leg orientation (Fig. 9G) (degrees): angle θ indicates the orientation of wetted midleg`s main axis (as well as the plane of the fan protracted under the leg, assuming fan surface`s plane includes the longitudinal axis of wetted leg) relative to the leg velocity vector. Angles closer to ‘90°’ indicate “more effective” employment of the full wetted midleg length in pushing the water surface dimple along the leg velocity vector and creating dimple asymmetry along the velocity vector. Under the assumption that longitudinal axis of wetted leg on the water surface approximately lies within the plane of the R. distincta fan`s surface protracted under water, θ values closer to 90° indicate that the angle between the fan surface and the fan movement direction is near perpendicular and hydrodynamic drag from the fan pushing backwards contributes to thrust.

    We additionally extracted three variables from the body movements :

    Body velocity (mm/s): distance (mm) traveled by the body center (position derived from the average of head and abdomen tip positions) between consecutive frames divided by the latency between the two consecutive frames (1/fps).

    Body acceleration (mm/s2): rate of change of body velocity derived from each pair of consecutive body velocity values divided by the latency between the two frames.

    Net force (µN): body acceleration (mm/s2) multiplied by insect body mass (mg) and by 0.001 for unit conversion. It represents a horizontal vector of net thrust force during a stroke.

    Fig. 9Graphical definitions of variables extracted from video recordings of thrust-generating strokes. (A) Digitized tracking points on the insect body and midleg. (B–D) Joint angles: (B) femur angle relative to the body axis, (C) femur-tibia angle, and (D) tibia-tarsus angle. (E–G) Indices of “effectiveness” for midleg orientation and movement: (E) effectiveness of leg velocity direction, (F) effectiveness of leg length`s use, (G) effectiveness of wetted leg orientation. Further details are provided in the ‘Methods’ section. (H) Side-by-side images of study species–Rhagovelia distincta and Gerris latiabdominis.Full size imageKinematic characterization of a strokeThe following kinematic variables were extracted from 21 strokes of six individuals of R. distincta and 12 strokes of six individuals of G. latiabdominis (1 value per stroke):

    Distance traveled by wetted midleg (mm): sum of frame-by-frame distances of wetted midleg midpoint was measured along the actual trajectory of the midpoint during the thrust phase of a stroke.

    Midleg`s stroke amplitude (mm): direct straight-line distance from the initial (beginning of thrust stroke) to the final (end of thrust stroke; when midleg`s velocity vector is no longer opposite to the body velocity vector) positions of wetted midleg midpoint was measured.

    Average angular velocity (degrees/s): mean of all frame-by-frame midleg angular velocities calculated within the thrust phase of a stroke.

    Average midleg velocity (mm/s): mean of all frame-by-frame leg velocities calculated within the thrust phase of a stroke.

    Leg velocity at maximum body acceleration (mm/s): leg velocity that corresponds to the maximum body acceleration (i.e., maximum horizontal net force) within the thrust phase of a stroke.

    Maximal body acceleration (mm/s2): maximum value of body acceleration within the thrust phase of a stroke.

    Maximal net thrust force (µN): maximum value of body force within the thrust phase of a stroke.

    Final body velocity (mm/s): final value of body velocity (between the last two consecutive frames) in the thrust phase of a stroke.

    Midleg thrust duration (ms): latency from the initiation of midleg thrust movements to the moment of their disengagement.

    Statistical analysesAll analyses were performed in R version 4.3.133. We used linear mixed-effects models (with “individual” as the random factor; “lme4” package34; “lmerTest” package35 to compare the effects of average leg velocity, stroke duration, and distance traveled by wetted midleg on the final body velocity between the two species (interaction with categorical variable “species”). However, these independent variables were correlated, and the statistical models would not allow proper evaluations of these effects. Therefore, we extracted principal components using functions fa.parallel and principal from the “psych” R package36 from the pooled data for both species considering all seven kinematic variables: body velocity, maximum acceleration, average angular leg velocity, average leg velocity, midleg`s stroke amplitude, distance traveled by wetted midleg, and stroke duration. We focused on the variables with loading values > 0.7537.

    Data availability

    All data generated or analyzed during this study are available in supplementary files.
    ReferencesHu, D. L., Chan, B. & Bush, J. W. M. The hydrodynamics of water Strider locomotion. Nature 424, 663–666 (2003).Article 
    ADS 
    PubMed 

    Google Scholar 
    Andersen, N. M. The Semiaquatic Bugs: Phylogeny, Adaptations, Biogeography, and Classification (Scandinavian Science, 1982).Crumière, A. J. J. et al. Diversity in morphology and locomotory behavior is associated with niche expansion in the semi-aquatic Bugs. Curr. Biol. 26, 3336–3342 (2016).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Khila, A., Abouheif, E. & Rowe, L. Comparative functional analyses of ultrabithorax reveal multiple steps and paths to diversification of legs in the adaptive radiation of semi-aquatic insects. Evolution 68, 2159–2170 (2014).PubMed 

    Google Scholar 
    Andersen, N. M. A comparative study of locomotion on the water surface in semiaquatic Bugs (Insecta, Hemiptera, Gerromorpha). Vidensk Medd Dan. Naturhist Foren. 139, 337–396 (1976).
    Google Scholar 
    Steinmann, T., Cribellier, A. & Casas, J. Singularity of the water Strider propulsion mechanisms. J. Fluid Mech. 915, A118 (2021).Article 
    ADS 
    MathSciNet 

    Google Scholar 
    Santos, M. E. et al. Taxon-restricted genes at the origin of a novel trait allowing access to a new environment. Science 358, 386–390 (2017).Article 
    ADS 
    PubMed 

    Google Scholar 
    Ortega-Jimenez, V. & Bhamla, S. Extreme maneuvering and hydrodynamics of Rhagovelia water striders. APS Div. Fluid Dyn. Meet. Abstr. E13-007 (2021). (2021).Ortega-Jimenez, V. et al. Ultrafast elastocapillary fans control agile maneuvering in ripple Bugs and robots. Science 389, 811–817 (2025).Article 
    ADS 
    PubMed 

    Google Scholar 
    Perez Goodwyn, P. J., Voigt, D. & Fujisaki, K. Skating and diving: changes in functional morphology of the setal and microtrichial cover during ontogenesis in Aquarius paludum Fabricius (Heteroptera, Gerridae). J. Morphol. 269, 734–744 (2008).Article 

    Google Scholar 
    Cho, Y. Gravity–capillary wave-making resistance on deep water. Phys. Fluids. 36, 017125 (2024).Article 
    ADS 

    Google Scholar 
    Kim, W. et al. Physics of sliding on water explains morphological and behavioral allometry across a wide range of body sizes in water striders (Gerridae). Proc. R. Soc. B 291, 20241357 (2024).Reaver, N. G. F. et al. Hydrodynamic controls on primary producer communities in spring-fed rivers. Geophys. Res. Lett. 46, 4715–4725 (2019).Article 
    ADS 

    Google Scholar 
    Bhavikatti, S. S. Mechanics of Solids (New Age International (P) Ltd., 2010).Zenkert, D. An Introduction To Sandwich Construction (Chameleon, 1995).Wu, X. & Zhu, Y. Gradient and lamellar heterostructures for superior mechanical properties. MRS Bull. 46, 244–249 (2021).Article 
    ADS 

    Google Scholar 
    Molnar, K., Vas, L. M. & Czigany, T. Strength of electrospun single nanofibers through modeling tensile behavior of the nanofibrous mat. Compos. Part. B Eng. 43, 15–21 (2012).Article 

    Google Scholar 
    Wingert, M. C. et al. Strong size-dependent stress relaxation in electrospun polymer nanofibers. J. Appl. Phys. 121, 015103 (2017).Article 
    ADS 

    Google Scholar 
    Zhang, X., Zhang, C. & Ma, P. Mechanical properties of Hollow polyester monofilament: compression and tension behaviors. J. Eng. Fibers Fabr. 14, 1–6 (2019).
    Google Scholar 
    Koehl, M. A. R. Hairy little legs: feeding, smelling and swimming at low Reynolds numbers. Contemp. Math. 141, 33–64 (1993).Article 

    Google Scholar 
    Lamont, E. I. & Emlet, R. B. Permanently fused setules create unusual folding fans used for swimming in cyprid larvae of barnacles. Biol. Bull. 235, 185–194 (2018).Article 
    PubMed 

    Google Scholar 
    Uesugi, K., Mayama, H. & Morshima, K. Proposal of a water-repellency model of water Strider and its verification by considering directly measured Strider leg-rowing force. J. Photopolym. Sci. Technol. 33, 185–192 (2020).Article 

    Google Scholar 
    Uesugi, K. Water-repellency model of the water Strider Aquarius paludum paludum by the curved structure of leg micro-hairs. J. Photopolym. Sci. Technol. 34, 393–399 (2021).Article 

    Google Scholar 
    Feng, X. Q. et al. Superior water repellency of water Strider legs with hierarchical structures: experiments and analysis. Langmuir 23, 4892–4896 (2007).Article 
    PubMed 

    Google Scholar 
    Gao, X. & Jiang, L. Water-repellent legs of water Striders. Nature 432, 36 (2004).Article 
    ADS 
    PubMed 

    Google Scholar 
    Koh, J. S. et al. Jumping on water: surface tension–dominated jumping of water Striders and robotic insects. Science 349, 517–521 (2015).Article 
    ADS 
    PubMed 

    Google Scholar 
    Yang, E. et al. Water Striders adjust leg movement speed to optimize takeoff velocity for their morphology. Nat. Commun. 7, 13698 (2016).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Fan, J., Zhang, W., Yuan, B. & Liu, G. Research on propulsion generation mechanism of frog swimming. Adv. Mech. Eng. 9, 1–8 (2017).Article 

    Google Scholar 
    Jackson, P. S. Performance prediction for olympic kayaks. J. Sports Sci. 13, 239–245 (1995).Article 
    PubMed 

    Google Scholar 
    Michael, J. S., Smith, R. & Rooney, K. B. Determinants of Kayak paddling performance. Sports Biomech. 8, 167–179 (2009).Article 
    PubMed 

    Google Scholar 
    Johansson, C. & Norberg, U. M. L. Lift-based paddling in diving Grebe. J. Exp. Biol. 204, 1687–1696 (2001).Article 
    PubMed 

    Google Scholar 
    Everitt, B. S. & Hothorn, T. A Handbook of Statistical Analyses Using R (Chapman & Hall/CRC, 2010).R Core Team. R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, 2023).Bates, D. et al. Parsimonious mixed models. Preprint at (2015). https://arxiv.org/abs/1506.04967Kuznetsova, A., Brockhoff, P. B. & Christensen, R. H. B. lmerTest package: tests in linear mixed effects models. J. Stat. Softw. 82, 1–26 (2017).Article 

    Google Scholar 
    Revelle, W. How To Use the Psych Package for Regression and Mediation Analysis (Northwestern University, 2023). https://personality-project.org/r/psych/Rutherford, G. S. W. et al. Multivariate data analysis with readings. Statistician 37, 484 (1988).Article 

    Google Scholar 
    Download referencesFundingThis work was supported by the National Research Foundation of Korea(NRF) grant funded by the Korea government(MSIT)(IRIS RS-2024-00343461; RS-2025-00514508); DGIST Start-up Fund Program nr 20200810 and individual mid-career grant 2022R1A2C1006090 of the Ministry of Science, ICT and Future Planning of Korea.Author informationAuthors and AffiliationsSchool of Biological Sciences, Seoul National University, Seoul, South KoreaSang Yun Bang, Woojoo Kim, Jeongseop Lee, Jinseok Park & Piotr Grzegorz JablonskiInsitutite of Biodiversity, Seoul National University, Seoul, South KoreaWoojoo KimResearch Institute of Basic Sciences, Seoul National University, Seoul, South KoreaWoojoo Kim & Jinseok ParkSoft Foundry Institute, College of Engineering, Seoul National University, Seoul, Republic of KoreaVersha KhareLaboratory of Integrative Animal Ecology, Department of New Biology, DGIST, Daegu, South KoreaSang-im LeeMuseum and Institute of Zoology, Polish Academy of Sciences, Warsaw, PolandPiotr Grzegorz JablonskiAuthorsSang Yun BangView author publicationsSearch author on:PubMed Google ScholarWoojoo KimView author publicationsSearch author on:PubMed Google ScholarJeongseop LeeView author publicationsSearch author on:PubMed Google ScholarJinseok ParkView author publicationsSearch author on:PubMed Google ScholarVersha KhareView author publicationsSearch author on:PubMed Google ScholarSang-im LeeView author publicationsSearch author on:PubMed Google ScholarPiotr Grzegorz JablonskiView author publicationsSearch author on:PubMed Google ScholarContributionsSY.B.: conceptualization, methodology, validation, formal analysis, investigation, data curation, writing main manuscript, writing-review & editing, and visualization; W.K.: formal analysis, investigation, writing-review & editing, and funding acquisition; J.L.: validation and investigation; J.P.: investigation, writing-review & editing, and funding acquisition; V.K.: investigation and funding acquisition; S-i.L.: conceptualization, resources, writing-review & editing, supervision and funding acquisition; P.G.J.: conceptualization, resources, writing-review & editing, and supervision.Corresponding authorsCorrespondence to
    Sang-im Lee or Piotr Grzegorz Jablonski.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 3Supplementary Material 4Supplementary Material 5Supplementary Material 6Supplementary Material 7Rights 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 articleBang, S.Y., Kim, W., Lee, J. et al. Distinct kinematics and micromorphology for symmetrical rowing and sliding on water in ripple bugs and water striders.
    Sci Rep 15, 44777 (2025). https://doi.org/10.1038/s41598-025-28453-0Download citationReceived: 16 June 2025Accepted: 11 November 2025Published: 29 December 2025Version of record: 29 December 2025DOI: https://doi.org/10.1038/s41598-025-28453-0Share 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

    Supplementary Material 4

    Supplementary Material 5

    Supplementary Material 6

    Supplementary Material 7 More

  • in

    Predicting natural enemy efficacy in biological control using ex-ante analyses

    AbstractMassive losses in agricultural and natural systems accrue globally due to invasive species, and yet the success rate of natural enemy introductions to control them is low. The high failure rate is due to the unknown efficacy of the introduced natural enemies. Furthermore, reviews of prior biological control efforts have not led to the development of assessment methods to predict their pre-release efficacy. To demonstrate a potential solution, we deconstructed the biological control of the invasive cassava mealybug (CM) and cassava green mite (CGM) in Africa using weather-driven metapopulation tri-trophic physiologically based demographic models (PBDMs). Bioeconomic analysis of the simulation results enabled parsing the contributions of the introduced natural enemies and endemic fungal pathogens to the control of CM and CGM and to the recovery of cassava yield across the vast ecological zones of Africa. The analysis shows that ex-ante pre-release analyses of natural enemy efficacy would have correctly predicted the biological control of the two pests. PBDM analyses of other biological control programs explained their success and/or failure. The results suggest well-parameterized mechanistic models can predict pre-release the efficacy of natural enemies and become an important instrument in increasing global food security.

    Similar content being viewed by others

    Machine learning and bifurcation analysis in a discrete predator-prey model with neem-induced mortality

    Article
    Open access
    19 November 2025

    A population based expression atlas provides insights into disease resistance and other physiological traits in cassava (Manihot esculenta Crantz)

    Article
    Open access
    07 December 2021

    Multidimensional β-diversity responses to global change: a meta-analysis highlighting divergent effects on plant communities

    Article
    Open access
    23 November 2025

    IntroductionMassive losses occur globally due to invasive species that may disrupt extant food webs in novel environments. Classical biological control has a proven record of success in controlling invasive pests, with the control of cottony cushion scale (Icerya purchasi Maskell) in California citrus in 1888 by the introduced vedalia beetle (Rodolia cardinalis (Mulsant)) and the parasitic fly Cryptochaetum iceryae (Williston) being the hallmark example1. Since the late nineteenth century, more than 2000 natural enemy species (agents) have been released in programs against approximately 400 invasive species worldwide2,3, resulting in partial or complete control of 226 invasive insect and 57 invasive weed species, with only ten cases resulting in negative outcomes4. However, the rate of introductions slowed after 1990 when the perspective in biological control changed focus from benefit to risk assessment, highlighting the need for conceptual models to guide benefit-risk decisions4. But there is little evidence that particular biological traits of natural enemies consistently predict the success of biological control, a conclusion attributed to insufficient data5. In general, biological control is significantly higher in single-species releases against insect pests than in multiple-species releases, but this contrasts with weeds, where control increases with the number of species released6.Theoretical differential models of classical biological control systems abound in the literature, with most concerning host-parasitoid interactions. These models have been used to make predictions about the independent effects of natural enemy attributes (e.g., sex ratio, reproductive strategies, fecundity, host feeding, developmental delays, dispersal rates, refuges, etc.) on system equilibrium with inference about control7. However, the predictions of such models are difficult to assess in the field, though some models have provided important theoretical insights8,9. The call to develop realistic field models to fill the gap between theory and practice has a long history7,10,11,12,11,, with some13 calling for an inclusive theory of biological control that incorporates the demographic and genetic processes to address the establishment and impact of introduced natural enemies.Field applications of models require incorporating the weather-driven, bottom-up effects of plants on higher trophic levels and the top-down effects of consumer behavior and physiology (sensu Hairston14). However, because of the myriad variant biology and complexity of species interactions, predicting the efficacy of natural enemy introductions has been assumed to be problematic10, but we demonstrate here that this need not be the case. Still, to determine what aspects of the biology of host-natural enemies are responsible for control or failure, we must be able to deconstruct holistically the weather-driven biology and dynamics of systems. Recent work demonstrated the ability to model the geospatial distribution and relative abundance of multiple invasive species15, but a critical gap remains: the ability to move from predicting where a pest or natural enemy might establish to predicting how well a specific biological control agent will perform before it is released. This study attempts to bridge some of this gap.In practice, we must use post hoc analyses of past natural enemy introductions to assess what aspects of the biology were crucial to control or failure. To illustrate this process, we deconstruct the highly successful biological control of two pests of cassava (Manihot esculenta Crantz) on the local scale and across the vast ecological zones of Africa using tri-trophic physiologically based demographic models (PBDMs) of the species in the system (Fig. 1A)16. First, we provide historical overview of cassava, the two exotic pests and the endemic and introduced natural enemies. A brief review of the biology of the species is given in the text, with greater details reported in the cited references and in the Supplemental Materials. We then simulate the weather-driven dynamics of the system across Africa and conduct marginal bioeconomic analyses of the area-wide results to parse the contributions of the natural enemies to the control of the two pests and to yield recovery. The approach is applicable to any system15 and demonstrating this is a thrust of our study. Self-reference is unavoidable as much of the relevant literature on the biological control of cassava pests in Africa and the development of the physiologically based model structure is by the authors.Historical perspective of cassava in AfricaCassava was introduced to Africa from Brazil during the slave trade in the sixteenth century, becoming an important staple crop across sub-Saharan Africa17. Two Neotropical pests, the cassava mealybug (Phenacoccus manihoti Matile-Ferrero (Hemiptera, Pseudococcidae)) (CM) and cassava green mite (Mononychellus tanajoa (Bondar), (Trombidiformes, Tetranychidae)) (CGM) were accidentally introduced to Africa early in the 1970s on cassava cuttings for use in plant breeding18,19. The two pests quickly spread across the African cassava zone causing massive yield losses creating severe food insecurity for more than 200 million Africans20.In Africa, CM was attacked by native generalist ladybird beetle predators in the genera Hyperaspis, Exochomus and Diomus, but they had minimal impact21,22. CM was also infected by the endemic fungal pathogen (Neozygites fumosa (Zygomycetes, Entomophthorales))23, especially during the rainy season24. Foreign exploration for effective natural enemies was conducted in the centers of cassava origin in Central and South America that resulted in the introduction, among others, of two hymenopterous parasitoid wasps (Anagyrus lopezi DeSantis and A. diversicornis Howard) (Hymenoptera, Encyrtidae)25. Both parasitoids were widely released, but A. diversicornis appears to have failed beyond the initial release sites in Benin26,27, while A. lopezi established widely and controlled CM populations at low levels across Africa28,29 and in Asia where CM subsequently spread30.CGM in Africa was attacked by native generalist insect predators and an endemic fungal pathogen, but with limited impact18. Eleven species of predators of CGM in the family Phytoseiidae were introduced from South America during the period 1984–2001, and more than 11.3 million were reared and released in 20 African countries31. Of these, only three of the predators established (Neoseiulus idaeus Denmark & Muma, Amblydromalus manihoti Moraes, and Typhlodromalus aripo De Leon), but only A. manihoti and T. aripo spread widely reducing CGM populations by half and increasing cassava yields by a third31. A virulent strain of the CGM-specific fungus Neozygites tanajoae was also introduced from Brazil to Benin, where local impact was observed32, but its spread was not monitored.Fig. 1A physiologically based demographic tri-trophic meta-population cassava system for Africa: (A) tri-trophic species interactions on individual cassava plants where single arrows indicate the direction of dry matter flow, double arrows indicate intraspecific competition, and pathogens are indicated as small symbols for conidia above CM and CGM, (B) the within plant flow of dry matter to age structured subunit populations, to herbivores (i.e., CM and CGM) and to higher tropic levels, and (C) individual plant interactions with each plant having populations of the species depicted in 1A with movement of mobile arthropod stages between plants governed by species-specific resource supply/demand ratios16. The map for Africa was developed using the open source GRASS GIS33 and open access geospatial data34.Full size imageResultsAs a guide, we first simulate the weather driven dynamics of the system components across Africa and then conduct bioeconomic analyses of the results to parse the contributions of the natural enemies to yield recovery and to the control of the two pests.CassavaThe prospective geographic distribution of average root yield per plant (grams) during the 1981–1990 period is mapped in Fig. 2A, while the same data masked for the historical distribution of cassava35 are mapped in Fig. 2B. Average annual degree days computed using a nonlinear model (dd > 14.86 °C) for cassava (Supplemental Materials) are mapped in Fig. 2C, and average annual rainfall is mapped in Fig. 2D with the limits of cassava production indicated by dashed lines35. The simulated distribution of cassava coincides well with the distribution of rainfall above the ~ 750 mm isohyet. The model predicts yields in South Sudan and parts of Uganda, Kenya, and Ethiopia (Fig. 2A vs. masked Fig. 2B), where alternative crops (e.g., maize, sorghum) are more prevalent. The gaps in Central Africa (Fig. 2B) are forested areas. All subsequent results are masked for the observed distribution of cassava35.Fig. 2Simulated pest-free cassava in Africa during years 1981–1990: (A) prospective distribution and average root yield (g dry matter per plant), (B) root yield data from Fig. 2A masked for the known distribution of cassava35 (public repository: https://doi.org/10.6084/m9.figshare.22491997), (C) average annual degree days > 14.85 °C for cassava estimated using a nonlinear model (Supplemental Materials) based on cassava growth rates data36, and (D) mean annual mm of rainfall with the limits of cassava circumscribed by the dashed white lines35.Full size imageImpact of cassava mealybug on yield—Absent natural control of CM, simulated average annual root yields (g dry matter per plant) over the 1981–1990 period are mapped in Fig. 3A. log10 average annual cumulative CM active stages per plant given the action of the endemic fungal pathogen are mapped in Fig. 3B. Prospective root yield losses are mapped in Fig. 3C as the difference of data in Figs. 2B and 3A. Model prediction of the geographic distribution of CM (Fig. 3B) agrees qualitatively with the distribution of CM estimated using the species distribution model CLIMEX37 (inset in Fig. 3B).Fig. 3Prospective per plant averages over the 1981–1990 period before biological control of CM given the effect of the endemic fungal pathogen: (A) prospective distribution of cassava root yield (g dry matter per plant) after the invasion of the cassava mealybug (CM), (B) prospective distribution of log10 cumulative annual mealybug active stages per plant with an inset clip of the predicted CLIMEX distribution of CM37, and (C) yield loss per plant due to uncontrolled mealybug computed as the difference of yields in Fig. 2B minus yields in Fig. 3A. The small difference in the scales between Figs. 2B and 3A is due to GIS system interpolation.Full size imageBiological control of cassava mealybug—Initial releases of the two parasitoids occurred at Cotonou, Benin, West Africa, and using Jun 30, 1982 to June 30, 1985 weather, rapid local control of CM by A. lopezi, A. diversicornis and the endemic fungal pathogen is predicted (Fig. 4A). Similar data for Ibadan, Nigeria are illustrated in the inset38. A. lopezi in concert with rainfall/fungal mortality suppress CM populations to very low levels with near total displacement of A. diversicornis16. However, absent A. lopezi, high populations of A. diversicornis are predicted but do not provide economic control (Fig. 4B). However, despite the minor role of A. diversicornis in suppressing CM locally27, it was included in the Africa-wide study to explore its potential in other ecological regions.Fig. 4Simulated average per plant population dynamics at Cotonou, Benin during 6/30/1982 to 6/30/1985: (A) cassava mealybug (CM), and CM parasitized by A. lopezi and A. diversicornis given the impact of fungal pathogen mortality on CM, and (B) CM and A. diversicornis parasitized CM given fungal mortality. Daily rainfall is indicated as the blue lines in 4A, and the inset shows data from Ibadan, Nigeria38.Full size imageProspective yields across the cassava belt rebounded to ~ 95% of pre-CM invasion levels (Figs. 3A vs. 5A) due principally to the control of CM by A. lopezi (Figs. 3B vs. 5B). The distribution and log10 average cumulative densities of A. lopezi and A. diversicornis post control of CM are mapped in Figs. 5C,D respectively showing the distribution of A. lopezi is wider and its densities considerably higher than that of A. diversicornis. High CM densities are predicted in drier northern and eastern areas where cassava production is marginal due to lower rainfall that reduces parasitoid and pathogen efficacy (see discussion).Fig. 5Prospective annual averages per plant during 1981–1990 after the release of A. lopezi and A. diversicornis: (A) relative root yield (g dry matter), (B) log10 cumulative CM active stages per year, (C) log10 cumulative A. lopezi parasitized mealybugs, and (D) log10 cumulative A. diversicornis parasitized mealybugs. The effect of the endemic fungal pathogen on CM is included in all sub figures, and all results are masked for the distribution of cassava35.Full size imageMarginal analysis of cassava yield with CM—Data for lattice cells with average root yield greater than 1500 g per plant (i.e., hereafter called the cassava belt) are used here and in all later analyses. The dependent variable Y in Eq. 1 is grams root dry matter per plant while the independent variables ((x_{{_{i} }}^{ + } = [0, , 1])) indicate absence or presence values for the species associated with specific Y values. The independent variables ((x_{{_{i} }}^{ + })) are mealybug (CM+), A. lopezi (Al+), A. diversicornis (Ad+), and pathogen (P+). Similar notation is used in all later analyses (see methods)$$begin{gathered} grams , root = 3464.8 – 1085CM^{ + } + 958.0Al^{ + } + 666.1Ad^{ + } + 378.2P^{ + } hfill \ quad quad quad quad quad – 687.9Ad^{ + } Al^{ + } – 260.5Ad^{ + } P^{ + } – 317.7Al^{ + } P^{ + } + 220.9Ad^{ + } Al^{ + } P^{ + } hfill \ quad quad quad quad quad R^{2} = 0.171, , F = 9,023.1, , df = 350,442 hfill \ mean , (t , value): , hfill \ CM^{ + } = , 0.860 , ( – 210.1), , Al^{ + } = { 0}{text{.533 }}(182.1), , Ad^{ + } = , 0.386 , (25.3), hfill \ P^{ + } = , 0.499 , (68.7), , Ad^{ + } Al^{ + } = , 0.266 , ( – 26.1), , Ad^{ + } P^{ + } = 0.253( – 9.6), , Al^{ + } P^{ + } = 0.268( – 43.3), hfill \ Ad^{ + } Al^{ + } P^{ + } = 0.132 , (8.1), hfill \ end{gathered}$$
    (1)
    Using average values for the (x_{{_{i} }}^{ + }), the marginal average root dry matter loss per plant due to CM is ({raise0.5exhbox{$scriptstyle {partial Y}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial CM^{ + } }$}}) = − 1085 g, while the average marginal contributions to yield recovery by Al+, Ad+ and P+ are ({raise0.5exhbox{$scriptstyle {partial Y}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial Al^{ + } }$}}) = 575.8 g, ({raise0.5exhbox{$scriptstyle {partial Y}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial Ad^{ + } }$}}) = 233.1 g, and ({raise0.5exhbox{$scriptstyle {partial Y}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial P^{ + } }$}}) = 153.8 g respectively for a total of 962.7 g. The positive contribution of A. diversicornis to yield occurs because Al+ was absent in about half of the simulations, and A. diversicornis partially suppresses CM (e.g., Fig. 4B). Given Al+, the expected contribution of Ad+ is low (Fig. 5D) as seen by excluding Ad+ from the analysis in Eq. 2.$$begin{gathered} grams , root = 3465.0 – 1091.0CM^{ + } + 955.0Al^{ + } + 596.7P^{ + } – 557.3Al^{ + } P^{ + } hfill \ quad quad quad quad quad R^{2} = 0.155, , F = 16,126, , df = 231,295 hfill \ mean , (t , value): , hfill \ quad quad quad quad quad CM^{ + } = 0.860 , ( – 209.39), , Al^{ + } = {0}{text{.533 }}(202.77), , P^{ + } = , 0.499 , (123.9), , Al^{ + } P^{ + } = 0.268( – 94.0) hfill \ end{gathered}$$
    (2)
    Using average values,({raise0.5exhbox{$scriptstyle {partial Y}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial Al^{ + } }$}}) = 676.9 g and ({raise0.5exhbox{$scriptstyle {partial Y}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial P^{ + } }$}}) = 299.7 g for a combined marginal compensation to yield of 976.0 g, which is roughly the same as in Eq. 1. The contribution of (Al^{ + }) to yield is ~ 2.25-fold that of (P^{ + }) supporting field observations28,29,39,40. The positive effect of (P^{ + }) are due to rainfall increasing both plant growth and fungal mortality to CM.Control of CM—Cumulative daily CM per plant per year (i.e., CM) is the dependent variable, and (Al^{ + }), (P^{ + }) are the independent variables in Eq. 3.$$begin{gathered} CM = {0}{text{.36616}}x10^{7} – 0.3370×10^{7} Al^{ + } + 0.3131×10^{6} P^{ + } – 0.4914×10^{6} Al^{ + } P^{ + } hfill \ quad quad quad quad quad R^{2} = 0.119, , F = {34,083}{text{.1}}, , df = 760,116 hfill \ mean , (t , value): , hfill \ quad quad quad quad quad Al^{ + } = , 0.518 , ( – 211.0), , P^{ + } = , 0.495 , (19.0), , Al^{ + } P^{ + } = 0.259 , (21.4) hfill \ end{gathered}$$
    (3)
    Using mean values, ({raise0.5exhbox{$scriptstyle {partial CM}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial Al^{ + } }$}} = – 3,613,246CM) and ({raise0.5exhbox{$scriptstyle {partial CM}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial P^{ + } }$}} = 58,555CM) showing the large suppressive effect of Al+ on CM, while the positive effect of P+ is the net effect of precipitation on pathogen mortality to CM and the enhancement of plant growth that increases CM. The combined action of Al+ and P+ reduces average cumulative CM active stages ~ 97%.Biological control of cassava green miteThe simulated dynamics of the cassava/CGM/predator system at the initial release site of Ikpinlè, Benin during 6/1/1990–6/30/2000 are summarized in Fig. 6. Before the introduction of the exotic predators, CGM mortality was low and due largely to the interaction of rainfall and an endemic fungal pathogen41,42 (Fig. 6A and B). After the introduction of the mite predators, CGM densities greatly declined due to the action T. aripo and less to A. manihoti, with highest CGM densities occurring during the dry season (Fig. 6C; see43). The simulation results confirm field observations31,44 that T. aripo is the most effective predator (Fig. 6D). Monthly estimates of CGM and predator densities on thirty plants during 1994–1997 at Ikpinlè, Benin summarized in Fig. 6E shows the persistence of the CGM-T. aripo interaction over multiple seasons.Fig. 6Simulated population dynamics of CGM and introduced predators per plant given the combined action of rainfall and endemic fungal pathogens using weather data for Ikpinlè, Benin during 6/1/1990 to 6/30/2000: (A) mm of daily rainfall, (B) CGM active stages absent predation, and (C) CGM with T. aripo and A. manihoti, (D) dynamics of T. aripo and A. manihoti active stages, and (E) field observations on 30 plants at approximately monthly intervals of average active stage counts of CGM (grey area) on the first developed leaf and T. aripo (orange line) in the shoot tips (see Supplemental Material for CGM and predator within plant distribution and behavior). The linear regression of data in Fig. 6E is (T. , aripo = 0.457CGM,{text{ R}}^{2} = 0.675) (Yaninek, unpublished data). The horizontal lines at 5000 in 6B and 6C are reference lines.Full size imageImpact of cassava green mite—Average annual rainfall during 1991–2000 is mapped in Fig. 7A, and prospective average cassava yields per plant absent CGM are mapped in Fig. 7B. Given CGM and the action of the endemic fungal pathogen, prospective cassava yields losses range from ~ 15% to ~ 67% (Fig. 7B vs. C) which agrees with measured yield losses of 13–80%45.Fig. 7Simulated cassava system dynamics during 1991–2000: (A) average annual rainfall (mm) with the dashed line indicating the northern ~ 800 mm rainfall isohyet for the period, (B) prospective root yield (g dry matter per plant) absent CGM, (C) root yield with CGM+ and endemic pathogen P+, (D) root yield with CGM+, P+, T. aripo (Ta+), and A. manihoti (Am+), and (E) CGM active stages with P+, (F) CGM active stages with P+, Ta+ and Am+, (G) T. aripo active stages given P+ and Am+, and (H) A. manihoti active stages given P+ and Ta+. The simulation data are masked for the distribution of cassava35.Full size imageThe introduction of the mite predators T. aripo and A. manihoti increased prospective root yields to ~ 95% of potential in many areas (Fig. 7B vs. D) due to control of CGM (Fig. 7E vs. F). Further, the prospective distribution of CGM (Fig. 7E) is comparable to that predicted by Parsa et al.46 using the species distribution MaxEnt algorithm47. The post introduction average densities of the two predators are summarized in Fig. 7G and H showing T. aripo has a wider distribution and its maximum densities are ~ 1.5–2.0 fold greater than A. manihoti.Marginal analysis of cassava yield with CGM—The simulated cassava belt yield data is the dependent variable and absence-presence values for green mite (CGM+), T. aripo (Ta+), A. manihoti (Am+), and the pathogen (P+) are the independent variable in Eq. 4.$$begin{gathered} grams , root = 3466.8 – 1345.0CGM^{ + } + 508.7Ta^{ + } + 289.0Am^{ + } + 488.8P^{ + } hfill \ quad quad quad quad quad – 287.5Ta^{ + } Am^{ + } P^{ + } hfill \ quad quad quad quad quad R^{2} = 0.224, , F = 15674.5, , df = 272,134 hfill \ mean , (t , value): , hfill \ quad quad quad quad quad CGM^{ + } = 0.816 , ( – 266.2), , Ta^{ + } = 0.449{ (}154.7),Am^{ + } = 0.439 , (87.0) hfill \ quad quad quad quad quad P^{ + } = 0.608 , (133.2), , Ta^{ + } Am^{ + } P^{ + } = 0.1581( – 58.8) hfill \ end{gathered}$$
    (4)
    Substituting mean values in Eq. 4, average yield loss is ~ 23%, and the marginal effects are:({raise0.5exhbox{$scriptstyle {partial Y}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial CGM^{ + } }$}} = – 1345.0g), while ({raise0.5exhbox{$scriptstyle {partial Y}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial P^{ + } }$}} = 432.1g), ({raise0.5exhbox{$scriptstyle {partial Y}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial Ta^{ + } }$}} = 432.0g) and ({raise0.5exhbox{$scriptstyle {partial Y}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial Am^{ + } }$}} = 210.5g) for a combined total of 1,074.6 g. This is an ~ 80% reduction in CGM damage (i.e., 1,074.6 g/1345 g). The large positive value for ({raise0.5exhbox{$scriptstyle {partial Y}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial P^{ + } }$}}) is the net effect of rain increasing yield and pathogen mortality on CGM. The negative interaction term Ta+Am+P+ is reflective of interspecific competition resulting in a ~ 1.3% (i.e., 45.5 g) reduction in average yield.Control of CGM—Using annual cumulative CGM eggs plus immature stages per plant as the dependent variable and Ta+, Am+, P+ as independent variables gives Eq. 5.$$begin{gathered} CGM = 80,381.0 , – 0.1042 times 10^{6} Ta^{ + } – 0.5160 times 10^{5} Am^{ + } + 0.1791 times 10^{6} P^{ + } hfill \ quad quad quad quad quad + 0.3105 times 10^{6} Ta^{ + } Am^{ + } – 0.2618 times 10^{6} Ta^{ + } Am^{ + } P^{ + } hfill \ quad quad quad quad quad R^{2} = 0.303, , F = 23,659.0, , df = 272,135 hfill \ mean , (t , value): , hfill \ quad quad quad quad quad Ta^{ + } = 0.450 , ( – 154.4), , Am^{ + } = 0.439 , ( – 75.6), , P^{ + } = 0.608 , (293.3), hfill \ quad quad quad quad quad Ta^{ + } Am^{ + } = 0.291 , (232.4), , Ta^{ + } Am^{ + } P^{ + } = 0.158 , ( – 283.8) hfill \ end{gathered}$$
    (5)
    The coefficients of Ta+, Am+ and Ta+Am+P+ are negative indicating they reduce CGM levels, while the interaction term Ta+Am+ is positive and is reflective of intraguild predation or competition. The coefficient for P+ is positive due to the net effect of rainfall favoring plant growth and the impact of pathogen mortality on CGM. Substituting average values in Eq. 5, the average annual cumulative CGM stages is 168,412 CGM per plant. The marginal effects of Ta+, Am+, P+ are ({raise0.5exhbox{$scriptstyle {partial CGM}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial Ta^{ + } }$}} = – 37,768 , CGM), ({raise0.5exhbox{$scriptstyle {partial CGM}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial Am^{ + } }$}} = , 16,496 , CGM), ({raise0.5exhbox{$scriptstyle {partial CGM}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial P^{ + } }$}} = , 102,916 , CGM) respectively.Removing P+ to assess the action of the two predators alone yields Eq. 6.$$begin{gathered} CGM = 285,830 , – 0.1274 times 10^{6} Ta^{ + } – 0.7035 times 10^{5} Am^{ + } + 0.6769 times 10^{5} Ta^{ + } Am^{ + } hfill \ quad quad quad quad quad R^{2} = 0.293, , F = 16,459, , df = 119,011 hfill \ mean , (t , value): , hfill \ quad quad quad quad quad Ta^{ + } = 0.576 , ( – 191.7), , Am^{ + } = 0.524 , ( – 101.4), , Ta^{ + } Am^{ + } = 0.288 , (74.3) hfill \ end{gathered}$$
    (6)
    The marginal impact of Ta+ (({raise0.5exhbox{$scriptstyle {partial CGM}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial Ta^{ + } }$}} = – 91,930.4 , CGM)) is ~ threefold that of Am+(({raise0.5exhbox{$scriptstyle {partial CGM}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial Am^{ + } }$}} = – 31,360.5 , CGM)) and combined both predators reduce prospective per plant CGM populations by ~ 43%. Substituting average values in Eq. 6, average annual cumulative CGM stages are 195,080 CGM vs. 168,412 CGM (Eq. 5), indicating an irreplaceable mortality from P+ of ~ 14%.DiscussionAnnual losses worldwide from invasive species since 1960 are estimated at US$ 1.13 trillion48 with annual losses of more than US$ 140 billion in the USA49, and US$ 28 billion in the European Union, with projected losses of US$148 billion by 204050. Loss of staple food crops (e.g., cassava, maize, millet, sorghum) and of veterinary and human health in Africa and other developing areas are large but not well documented51,52. Classical biological control of invasive species using introduced natural enemies53 is a common practice, and when successful, control is self-sustaining unless disrupted. An exemplar modern biological control success was of the highly destructive cassava mealybug (CM) and cassava green mite (CGM) in Africa54 that kept food insecurity at bay for more than 200 million people40. Depending on monetary scenarios, the economic gains from the control of CM alone ranged from US$ 10–30 billion, with benefit–cost ratios of ~ 20055 and ~ 370–74056,57 that continue to increase. In contrast, the cost of the biological control program for CM and CGM was ~ US$16 million or about eight US cents per person affected20.But classical biological control programs have a high failure rate2,3 largely because natural enemies are released after host specificity screening without sufficient knowledge of their potential efficacy. Furthermore, analyses of natural enemy attributes in simple models fail to inform efficacy in specific cases5,7. A major contributing factor in the failures is that target invasive species may have wide native ranges with numerous strains adapted to local conditions58, with natural enemies adapted to them. Because the initial invasion of a novel area may come from any part of the range, finding the most efficacious strains of natural enemies often proves vexing. A documented example is the walnut aphid (Chromaphis juglandicola Kaltenbach) where the temperature dependent vital rates of the parasitoid Trioxys pallidus (Haliday, 1833) introduced from France were ill suited to the hot environmental conditions of the Central Valley of California, while an Iranian strain provided excellent control59. Physiological incompatibility of a parasitoid to its host may also occur as found in the alfalfa weevil (Hypera postica (Gyllenhal))60.The biology matters, and including the relevant complexity in weather-driven models is not difficult. By way of illustration, we show that an ex-ante analysis using our cassava system model and weather data from the invaded area would have given the correct predictions concerning the efficacy of the natural enemies for both CM and CGM across Africa. Specifically, the analyses capture astonishingly well the observed distribution and relative abundance of cassava, and of CM and CGM and their natural enemies, and their effects on cassava yield across Africa; all independent of species distribution records. The model captured the role of the biology, behavior, and interactions of the parasitoids A. lopezi and A. diversicornis in the control of CM (see21,26) and showed that natural enemies, rather than weather or resistant varieties, caused the decline of CM populations. The analysis explains why A. lopezi introduced from a small region in the Rio de la Plata Valley in South America quickly established with immediate high impact on CM across the ecological zones of Africa27,57. However, the analysis also shows that if only A. diversicornis had been introduced, partial control of CM would have resulted in failure (see Figs. 4 and 5).The search for predators of CGM was focused in Brazil using a climate matching program developed at CGIAR/CIAT in Colombia61. Though eleven species of mite predator were introduced to Africa, only three species became established, with A. manihoti and T. aripo spreading widely and T. aripo becoming the dominant predator in the field62. This occurred despite A. manihoti having higher demographic parameters. The analysis explains that the capacity of T. aripo to use plant resources (e.g., pollen and likely exudates) as an alternative food source was responsible for its success (see Supplemental Materials). Hence, if only A. manihoti had been introduced, only partial control of CGM would have occurred, resulting in failure. Furthermore, the model predicts, as observed28, that populations of CM (Fig. 5B) and CGM (Fig. 7F) may still develop in areas with marginal rainfall as mortality from fungal pathogens decrease.Complex food webs abound in agroecological and natural systems with myriad interactions across trophic levels63, and the entry of an invasive species at any trophic level may disrupt trophic interactions and seemingly present complexity beyond the capacity of PBDM analysis. But systems like alfalfa (Medicago sativa L.) have numerous food webs, and when the invasive spotted alfalfa aphid (SAA, Therioaphis trifolii (Monell)) invaded California, it severely disrupted the system regionally. A PBDM analysis of the natural enemies introduced for control of SAA enabled parsing the interacting but differing contributions of three parasitoids, a fungal pathogen and native coccinellid predators to the control of SAA across the different ecological zones where alfalfa grows. With control, the system recovered10, but this might not be the case in systems where critical components of food webs are irreparably changed or destroyed.Well-documented PBDM analyses of natural enemy efficacy include the coffee (Coffea arabica L.) system. The analysis explained why several introduced parasitoids of coffee berry borer (Hypothenemus hampei (Ferrari)) in the Americas fail to control the pest, but also enabled exploration of alternative effective harvest and cleanup control strategies64. Using a simple model, Abram et al.65 posited that egg parasitoids of the invasive Asian brown marmorated stink bug (Halyomorpha halys Stål) were unlikely to control it, while a tri-trophic PBDM system explained why, and further posited ex-ante that despite low levels of parasitism, tachinid parasitoids of late nymphal and of adult stages would augment egg parasitism and have an important role in suppressing the pest across its vast Palearctic range66. Similarly, only partial control of yellow starthistle (YST; Centaurea solstitialis L.) in North America occurred despite extensive releases of several seed head insects that insufficiently reduced seed survival. An ex-ante PBDM analysis of the YST system posits that control would be enhanced by the introduction of the rosette weevil (Ceratapion basicorne (Illiger)) that kills whole plants and further reduces seed production in survivors67. These examples and others summarized in the Supplemental Materials in Gutierrez et al.15 strongly suggest that ex-ante in silico analyses of natural enemy efficacy with well parameterized mechanistic models such as PBDMs using weather data from invaded regions would be strategic and cost-effective and fill a critical need enhancing global food security.Nevertheless, the physiological and behavioral parameters of the interacting species determine success or failure, and they vary widely. Hence, the biology matters and capturing it for each species in a food chain or web is crucial in predicting efficacy—no shortcut rules are apparent, but general methods for assembling PBDMs are increasingly available15. Ex-ante analyses of natural enemies in their native environments would further streamline field-based pre-release risk assessments68,69, with species distribution70 and genomic analyses71 being complementary. Such holistic analyses would further lower risk in natural enemy introduction4.While PBDM development is perceived as difficult72, the math is simple, and the process is quite straightforward15,73. The bottleneck is often finding appropriate data in the literature to parameterize PBDMs and/or the inability to develop the biological data in house—especially in the pest’s native environments. Further, studies on species biology are often deemed mere technical work and are not fully appreciated or rewarded professionally, and yet such data are key to answering larger ecological/economic/social questions (see74). To facilitate progress in PBDM development, Python-based software is being developed to enable non-experts to develop PBDMs75, and appropriate weather files to drive the models are increasingly available. Textbooks outlining the mathematics are available14,76.Among the myriad pest problems requiring holistic agro-eco-social analyses76 is whitefly (Bemisia tabaci (Glenn.)) that vectors mosaic disease and brown streak disease that currently constrain cassava production in Africa. Control of B. tabaci in resource-poor African cassava requires control methods that are low-cost and effective, such as traditional host-plant resistance and biological control—methods that are sustainable and readily disseminated77. The PBDMs for cassava and for the complicated biology of whitefly parasitoids78 could be used as a starting point to deconstruct this problem.As epilogue, the PBDM approach tends toward reductionism, and from a theoretical standpoint has roots in thermodynamics with an explicative basis for population interactions provided by bioeconomic supply–demand principles common to all economies including those of humans79–81,80,. This framework may provide a mechanistic pathway for ex-ante assessments of a natural enemy efficacy that can dramatically increase the return on investment in biological control and empower its use as an important tool in sustainable, climate-resilient agriculture; to move classical biological control beyond ‘release-and-hope for the best stage’—to increase ecological stability in agroecological and natural systems with greater foresight and precision. This was the raison d’être of our study, and while we caution that no model can be a one to one description of nature, we recognize that—as stated by Heisenberg82, cited from 83—”what we observe is not nature itself, but nature exposed to our methods of questioning”.MethodsPBDMs are time-varying life tables10,12,80 that capture the weather-driven daily development and temporal dynamics of species independent of time and place, a property that enables their use in climate change and GIS contexts. Weather-driven PBDMs of the cassava subsystems (Fig. 1A) linked by dry matter and energy flow to age-structured populations of plant subunits (Fig. 1B) were used in our study (80, see84). Cock et al.85 (see86) report a simulation model that predicts cassava growth and dry matter allocation to plant subunits but lacks demographic structure, limiting linkages to higher trophic levels. The cassava system model used is a meta-population composed of a population of plants (100 maximum, Fig. 1C) with each potentially having the full complement of interacting arthropod populations16 in the thousands of lattice cell across vast areas and ecological zones of Africa. As implemented, the PBDMs capture the weather driven, time varying bio-economics (supply/demand relationships) of species resource acquisition and allocation79,81.Fig. 8PBDM submodels: (A) schema for bio-economics of energy flow in a linear tri-trophic system (see text for allocation priorities), and (B) stylized biodemographic functions (BDFs) used to develop PBDMs: [(Ba) is the rate of development on temperature, (Bb) is the functional response for resource acquisition, (Bc) the age-specific reproductive profile at the optimum temperature, (Bd) the temperature scalar to correct reproduction from the optimum temperature, and (Be) is the mortality rate per day on temperature]. Note that similar shaped scalars to (Bd) and (Be) can be developed for the effects of relative humidity and other factors that in practice are not symmetrical.Full size imageCassava variety IITA TMS30572 is used as a generic prototype with movement of arthropods between plants modeled as a function of species-specific resource supply–demand ratios on each plant. The only change to our model16 is the use of a nonlinear developmental rate data for cassava36 that captures the limiting effects of high temperature. The cassava system model is modular, and any combination of species with links to cassava and as appropriate to each other can be introduced in the computations using Boolean variables. The system model was coded in Borland Delphi Pascal 3. The two major components of PBDMs are: the age/mass structured dynamics model and the parameterized biological functions capturing the biology of each species.Population dynamics model—The dynamics model accommodates age-mass structure and temperature-dependent developmental rates, and considerable aspects of the biology and behavior. Cohorts of ectotherm organisms have distributed maturation times with means and variance, and each stage/species may have different responses to temperature, resulting in different time scales for each. To accommodate the developmental biology and the behavior and physiology of the species, we use the time-invariant distributed maturation time dynamics model with attrition87 (see Supplemental materials), but the time-varying form is also available88. Other dynamics models may also be used12,89,90. Two approaches are used to parameterize the dynamics models: the metabolic pool (MP, Fig. 8A) and the biodemographic function approach (BDF, Fig. 8B)15.PBDM/MP models capture per capita biomass/energy acquisition and allocation to wastage, respiration, growth and reserves if immature and to reproduction in adults (Fig. 8A), and the effects of mortality on plant and arthropod populations. Plants search for light, water, and nutrients to produce photosynthate that is allocated to plant subunit vegetative growth and reproduction91. The PBDM/MP approach was first used to develop a cotton canopy model composed of linked age-mass structured populations of leaves, stem, root, and fruit92,93, and later to develop models of the growth, development, and reproduction of pest arthropods11.In PBDM/BDF models, the rates of development, births, and deaths of the species are estimated from age-specific life table studies conducted under an array of temperatures and other conditions73, but field studies can also be used94. The vital rates are the outcomes over the life cycle of a cohort of organisms of how they acquire and allocate energy, survive, and reproduce under the experimental conditions—i.e., the results of metabolic pool processes under the experimental conditions. BDFs summarize the effects of abiotic variables on species developmental, birth, and death rates across environmental conditions, and are used to parameterize the dynamics models. Common BDFs are depicted in a stylized manner in Fig. 8B, but other BDFs may be developed to accommodate additional aspects of the biology of species (e.g., the response to relative humidity; see Supplemental Materials). Temperature-dependent development (Fig. 8Ba) is used in both MP and BDF approaches to estimate daily increments of species-specific physiological time and age. The migration rates of arthropods between plants in both MP and BDF approaches are functions of species-specific resource acquisition success (i.e., 0 < supply/demand < 1).These methods were used to model our cassava system, noting that most of the species models are MP-based.Weather data—The system model is run using daily maximum and minimum temperature, precipitation, solar radiation, and relative humidity as inputs (see Supplemental Materials). Daily weather data for 40,691 ~ 25 × 25 km georeferenced lattice cells across Africa for the 1980–2010 period were sourced from the open access AgMERRA global weather dataset created by the Agricultural Model Intercomparison and Improvement Project (AgMIP, https://agmip.org/)95. The AgMERRA data were accessed through the Goddard Institute for Space Studies (GISS) of the National Aeronautics and Space Administration (NASA, https://data.giss.nasa.gov/impacts/agmipcf/). From the perspective of ectotherms, climate change is simply another weather pattern, but the effects of climate change on the system are beyond the scope of this study (see examples15).Simulations—The time step in the model is a day in physiological time units that differ daily for each species. The computation time for the daily dynamics of a 100-plant system per year at one location is ~ 6–10 s on a laptop computer, making the total time over an eleven-year period across 40,691 lattice cells inordinately long. Hence for feasible computations on a laptop computer, ten randomly spaced plants in 10,172 lattice cells in alternating latitude–longitude were used. This subset reduced the computation time considerably and yet provided sufficient grain to characterize the distribution and relative abundance of the species across Africa.The biological control efforts against CM occurred during 1980–1990 and during 1990–2000 for CGM, and hence all combinations of the two subsystems were run separately for the two time periods. We note, however, all species can be run concurrently. The same nominal initial conditions for species were used for all lattice cells, and the species absence-presence values (i.e., [0, 1]) and selected annual georeferenced summary simulation variables for all populations were written to text files. Because the system was assumed equilibrating to lattice cell weather, data from the first year were excluded in computing lattice cell means, standard deviations, and coefficients of variation of the annual summary variables. Only means of summary variables are used in GIS mapping and the bioeconomic analyses.GIS mapping—The open source GIS software GRASS96,97 (http://grass.osgeo.org/) was used to map the summary output data. The map of the current distribution of cassava cultivation in Africa35 is used in our analyses for comparative purposes. Other GIS data mapping layers were sourced from the public domain repository Natural Earth (https://www.naturalearthdata.com/). Inverse distance weighting bicubic spline interpolation was used to create a continuous raster surface of the simulation data.Bioeconomic analysis—A general binomial multiple linear regression model (Eq. 7) was used to summarize the Africa-wide summary of annual simulation data. Root dry matter yield (grams) or pest density (CM, CGM) are the dependent variables (Y) and species absence-presence values are the I independent dummy variables (i.e., (x_{{_{i} }}^{ + }), [0, 1]). Specifically,$$Y = f(x_{1} , , …..,x_{I} ) = a + sumlimits_{i = 1}^{I} {b_{i} x_{{_{i} }}^{ + } } + sumlimits_{j = 1}^{J} {c_{j} {}_{n}^{I} x_{j}^{ + } } + dprodlimits_{i = 1}^{I} {x_{i}^{ + } } + U$$
    (7)
    where ({}_{n}^{I} x_{j}^{ + }) is the jth interaction of the (x_{{_{i} }}^{ + }) taken n at a time, (prodlimits_{i = 1}^{I} {x_{{_{i} }}^{ + } }) is the product of the I variables, and U is the error term. Only significant regression terms (p ≤ 0.05) were included in the final model. The goal of the analyses was to estimate the marginal impact of species presence (x_{{_{i} }}^{ + }) on cassava yield (g dry matter) and pest density estimated as ({raise0.5exhbox{$scriptstyle {partial Yield}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial x_{i}^{ + } }$}}),({raise0.5exhbox{$scriptstyle {partial CM}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial x_{i}^{ + } }$}}) and ({raise0.5exhbox{$scriptstyle {partial CGM}$} kern-0.1em/kern-0.15em lower0.25exhbox{$scriptstyle {partial x_{i}^{ + } }$}}) given the average effects of the other independent variables.

    Data availability

    All the data used are available in the cited references, with the fitted BDF functions given in the supplemental materials. All the GRASS GIS geospatial data layers used in this analysis are available open access on Zenodo34 at https://doi.org/10.5281/zenodo.13254494.
    Code availability

    The Borland Delphi Pascal 3 code for the cassava tri-trophic PBDM system is available open source on Zenodo98 at https://zenodo.org/records/17559583 (associated GitHub repository: https://github.com/casasglobal-org/pascal-pbdm-cassava). The code for the geospatial software based on GRASS GIS used for mapping and geospatial analysis is available open source on Zenodo33 at https://zenodo.org/records/17551910 (associated GitHub repository: https://github.com/casasglobal-org/casas-gis).
    ReferencesQuezada, J. R. & DeBach, P. Bioecological and population studies of the cottony scale, Icerya purchasi Mask., and its natural enemies, Rodolia cardinalis Mul. and Cryptochaetum iceryae Wil., in southern California. Hilgardia 41, 631–688 (1973).Article 

    Google Scholar 
    Van Lenteren, J. C., Bale, J., Bigler, F., Hokkanen, H. M. T. & Loomans, A. J. M. Assessing risks of releasing exotic biological control agents of arthropod pests. Annu. Rev. Entomol. 51, 609–634 (2006).Article 
    PubMed 

    Google Scholar 
    Cock, M. J. W. et al. Trends in the classical biological control of insect pests by insects: An update of the BIOCAT database. Biocontrol 61, 349–363 (2016).Article 

    Google Scholar 
    Heimpel, G. E. & Cock, M. J. W. Shifting paradigms in the history of classical biological control. Biocontrol 63, 27–37 (2018).Article 

    Google Scholar 
    Segoli, M. et al. Trait-based approaches to predicting biological control success: Challenges and prospects. Trends Ecol. Evol. 38, 802–811 (2023).Article 
    PubMed 

    Google Scholar 
    Denoth, M., Frid, L. & Myers, J. H. Multiple agents in biological control: Improving the odds?. Biol. Control 24, 20–30 (2002).Article 

    Google Scholar 
    Mills, N. Biological control: The need for realistic models and experimental approaches. In Parasitoid Population Biology (eds Hochberg, M. E. & Ives, A. R.) 217–234 (Princeton University Press, 2000).Chapter 

    Google Scholar 
    Hawkins, B. A. & Cornell, H. V. Theoretical approaches to biological control (Cambridge University Press, 1999).Book 

    Google Scholar 
    Schreiber, S. J., Mills, N. J. & Gutierrez, A. P. Host-limited dynamics of autoparasitoids. J. Theor. Biol. 212, 141–153 (2001).Article 
    ADS 
    PubMed 

    Google Scholar 
    Gilbert, N., Gutierrez, A. P., Frazer, B. D. & Jones, R. E. Ecological relationships (W.H. Freeman and Co., 1976).
    Google Scholar 
    Gutierrez, A. P. & Baumgärtner, J. U. Multitrophic level models of predator-prey energetics: I. Age-specific energetics models – pea aphid Acyrthosiphon pisum (Homoptera: Aphididae) as an example. Can. Entomol. 116, 924–932 (1984).Article 

    Google Scholar 
    Gutierrez, A. P. Applied population ecology: A supply-demand approach (John Wiley and Sons, 1996).
    Google Scholar 
    Mills, N. J. An alternative perspective for the theory of biological control. Insects 9, 131 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hairston, N. G., Smith, F. E. & Slobodkin, L. B. Community structure, population control, and competition. Am. Nat. 94, 421–425 (1960).Article 

    Google Scholar 
    Gutierrez, A. P. et al. Geospatial physiologically based demographic modeling and analysis of thirteen invasive species in Africa: Why the biology matters. Environ. Sci. Eur. 37, 67 (2025).Article 

    Google Scholar 
    Gutierrez, A. P., Yaninek, J. S., Neuenschwander, P. & Ellis, C. K. A physiologically-based tritrophic metapopulation model of the African cassava food web. Ecol. Model. 123, 225–242 (1999).Article 

    Google Scholar 
    Ferguson, M. E., Shah, T., Kulakow, P. & Ceballos, H. A global overview of cassava genetic diversity. PLoS ONE 14, e0224763 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Yaninek, J. S. & Herren, H. R. Introduction and spread of the cassava green mite, Mononychellus tanajoa (Bondar) (Acari: Tetranychidae), an exotic pest in Africa and the search for appropriate control methods: a review. Bull. Entomol. Res. 78, 1–13 (1988).Article 

    Google Scholar 
    Matile-Ferrero, D. Une cochenille nouvelle nuisible au manioc en Afrique équatoriale, Phenacoccus manihoti N. Sp. [Homoptera, Coccoidea, Pseudococcidae]. Ann. Soc. Entomol. Fr. 13, 145–152 (1977).Article 

    Google Scholar 
    Herren, H. R. & Neuenschwander, P. Biological control of cassava pests in Africa. Annu. Rev. Entomol. 36, 257–283 (1991).Article 

    Google Scholar 
    Gutierrez, A. P., Neuenschwander, P. & Van Alphen, J. J. Factors affecting biological control of cassava mealybug by exotic parasitoids: a ratio-dependent supply-demand driven model. J. Appl. Ecol. 30, 706–721 (1993).Article 

    Google Scholar 
    Gutierrez, A. P. et al. An overview of a systems model of cassava and cassava pests in Africa. Insect Sci. Appl. 8, 919–924 (1987).
    Google Scholar 
    Le Rü, B. Etude de l’évolution d’une mycose à Neozygites fumosa (Zygomycetes, Entomophthorales) dans une population de la cochenille du manioc, Phenacoccus manihoti (Hom.: Pseudococcidae). Entomophaga 31, 79–89 (1986).Article 

    Google Scholar 
    Schulthess, F., Baumgärtner, J. U. & Herren, H. R. Factors influencing the life table statistics of the cassava mealybug Phenacoccus manihoti. Insect Sci. Applic. 8, 851–856 (1987).
    Google Scholar 
    Herren, H. R., Neuenschwander, P., Hennessey, R. D. & Hammond, W. N. O. Introduction and dispersal of Epidinocarsis lopezi (Hym., Encyrtidae), an exotic parasitoid of the cassava mealybug, Phenacoccus manihoti (Hom., Pseudococcidae), in Africa. Agric., Ecosyst. Environ. 19, 131–144 (1987).Article 

    Google Scholar 
    Alphen, J. V., Kraaijeveld, A. R. & Xu, C. R. A comparison of Epidinocarsis lopezi and E. diversicornis: A possible explanation for the failed introduction of E. diversicornis against cassava mealybug Phenacoccus manihoti into Africa. Mededelingen van de Faculteit Landbouwwetenschappen, Rijksuniversiteit Gent 55, 276–287 (1990).
    Google Scholar 
    Neuenschwander, P. Biological control of the cassava mealybug in Africa: A review. Biol. Control 21, 214–229 (2001).Article 

    Google Scholar 
    Neuenschwander, P. et al. Biological control of the cassava mealybug, Phenacoccus manihoti (Hom., Pseudococcidae) by Epidinocarsis lopezi (Hym., Encyrtidae) in West Africa, as influenced by climate and soil. Agric., Ecosyst. Environ. 32, 39–55 (1990).Article 

    Google Scholar 
    Neuenschwander, P. et al. Biological control of the cassava mealybug Phenacoccus manihoti (Hom., Pseudococcidae) by Epidinocarsis lopezi (Hym., Encyrtidae) in Malawi. Biocontrol Sci. Technol. 1, 297–310 (1991).Article 

    Google Scholar 
    Wyckhuys, K. A. G. et al. Continental-scale suppression of an invasive pest by a host-specific parasitoid underlines both environmental and economic benefits of arthropod biological control. PeerJ 6, e5796 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Yaninek, J. S. Biological control of the cassava green mite in Africa: overcoming challenges to implementation. In Biological control: a global perspective (eds Vincent, C., Goettel, M. S. & Lazarovits, G.) 28–37 (CAB International Publisher, UK, 2007). https://doi.org/10.1079/9781845932657.0028.Hountondji, F. C. C., Lomer, C. J., Hanna, R., Cherry, A. J. & Dara, S. K. Field evaluation of brazilian isolates of Neozygites floridana (Entomophthorales: Neozygitaceae) for the microbial control of cassava green mite in Benin, West Africa. Biocontrol Sci. Technol. 12, 361–370 (2002).Article 

    Google Scholar 
    Ponti, L., Gutierrez, A. P. & Neteler, M. CASAS GIS – Geospatial software based on GRASS GIS. Zenodo https://doi.org/10.5281/zenodo.17551910 (2025).Ponti, L., Gutierrez, A. P. & Neteler, M. GRASS GIS database for CASAS-PBDM (www.casasglobal.org) geospatial mapping and analysis. Zenodo https://doi.org/10.5281/zenodo.13254494 (2024).Tang, F. H. M. et al. CROPGRIDS: a global geo-referenced dataset of 173 crops. Sci Data 11, 413 (2024).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Keating, B. A. & Evenson, J. P. Effect of soil temperature on sprouting and sprout elongation of stem cuttings of cassava (Manihot esculenta Crantz.). Field Crops Res. 2, 241–251 (1979).Article 

    Google Scholar 
    Yonow, T., Kriticos, D. J. & Ota, N. The potential distribution of cassava mealybug (Phenacoccus manihoti), a threat to food security for the poor. PLoS ONE 12, e0173265 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Gutierrez, A. P. et al. Analysis of biological control of cassava pests in Africa. II. Cassava mealybug Phenacoccus manihoti. J. Appl. Ecol. 25, 921–940 (1988).Article 

    Google Scholar 
    Hennessey, R. D., Neuenschwander, P. & Muaka, T. Spread and current distribution of the cassava mealybug, Phenacoccus manihoti (Homoptera: Pseudococcidae) in Zaire. Int. J. Pest Manag. https://doi.org/10.1080/09670879009371451 (1990).Article 

    Google Scholar 
    Neuenschwander, P. et al. Impact assessment of the biological control of the cassava mealybug, Phenacoccus manihoti Matile-Ferrero (Hemiptera: Pseudococcidae), by the introduced parasitoid Epidinocarsis lopezi (De Santis) (Hymenoptera: Encyrtidae). Bull. Entomol. Res. 79, 579–594 (1989).Article 

    Google Scholar 
    Yaninek, J. S., Herren, H. R. & Gutierrez, A. P. Dynamics of Mononychellus tanajoa (Acari: Tetranychidae) in Africa: Seasonal factors affecting phenology and abundance. Environ. Entomol. 18, 625–632 (1989).Article 

    Google Scholar 
    Yaninek, J. S., Saizonou, S., Onzo, A., Zannou, I. & Gnanvossou, D. Seasonal and habitat variability in the fungal pathogens, Neozygites cf. floridana and Hirsutella thompsonii, associated with cassava mites in Benin, West Africa. Biocontrol Sci. Technol. 6, 23–34 (1996).Article 

    Google Scholar 
    Yaninek, J. S., Herren, H. R. & Gutierrez, A. P. The biological basis for the seasonal outbreak of cassava green mites in Africa. Insect Sci. Appl. 8, 861–865 (1987).
    Google Scholar 
    Yaninek, J. S. et al. Establishment and spread of Typhlodromalus manihoti (Acari: Phytoseiidae), an introduced phytoseiid predator of Mononychellus tanajoa (Acari: Tetranychidae) in Africa. Environ. Entomol. 27, 1496–1505 (1998).Article 

    Google Scholar 
    Yaninek, J. S., Gutierrez, A. P. & Herren, H. R. Dynamics of Mononychellus tanajoa (Acari: Tetranychidae) in Africa: effects on dry matter production and allocation in cassava. Environ. Entomol. 19, 1767–1772 (1990).Article 

    Google Scholar 
    Parsa, S. et al. Potential geographic distribution of two invasive cassava green mites. Exp. Appl. Acarol. 65, 195–204 (2015).Article 
    PubMed 

    Google Scholar 
    Phillips, S. J. & Dudik, M. Modeling of species distributions with Maxent: new extensions and a comprehensive evaluation. Ecography 31, 161–175 (2008).Article 
    ADS 

    Google Scholar 
    Cuthbert, R. N. et al. Biological invasion costs reveal insufficient proactive management worldwide. Sci. Total Environ. 819, 153404 (2022).Article 
    PubMed 

    Google Scholar 
    Pimentel, D., Lach, L., Zuniga, R. & Morrison, D. Environmental and economic costs of nonindigenous species in the United States. Bioscience 50, 53–65 (2000).Article 

    Google Scholar 
    Henry, M. et al. Unveiling the hidden economic toll of biological invasions in the European Union. Environ. Sci. Eur. 35, 43 (2023).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Pyšek, P. et al. Geographical and taxonomic biases in invasion ecology. Trends Ecol. Evol. 23, 237–244 (2008).Article 
    PubMed 

    Google Scholar 
    Khuroo, A. A., Reshi, Z. A., Rashid, I. & Dar, G. H. Towards an integrated research framework and policy agenda on biological invasions in the developing world: A case-study of India. Environ. Res. 111, 999–1006 (2011).Article 
    PubMed 

    Google Scholar 
    Huffaker, C. B. & Messenger, P. S. Theory and Practice of Biological Control. (Academic Press, San Francisco, USA, 1976). https://doi.org/10.1016/B978-0-12-360350-0.X5001-6.Neuenschwander, P., Borgemeister, C., Groote, H. D., Sæthre, M.-G. & Tamò, M. Perspective article: Food security in tropical Africa through climate-smart plant health management. Heliyon 9, e15116 (2023).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Norgaard, R. B. The biological control of cassava mealybug in Africa. Am. J. Agr. Econ. 70, 366–371 (1988).Article 

    Google Scholar 
    Zeddies, J., Schaab, R. P., Neuenschwander, P. & Herren, H. R. Economics of biological control of cassava mealybug in Africa. Agric. Econ. 24, 209–219 (2001).Article 

    Google Scholar 
    Neuenschwander, P. Harnessing nature in Africa. Nature 432, 801–802 (2004).Article 
    ADS 
    PubMed 

    Google Scholar 
    Gutierrez, A. P. et al. Stabilizing adaptation in an invasive species: Alfalfa weevil as a case study. Agric. For. Entomol. https://doi.org/10.1111/afe.12686 (2025).Article 

    Google Scholar 
    van den Bosch, R. et al. Biological control of the walnut aphid in California: Impact of the parasite. Trioxys pallidus. Hilg 47, 1–13 (1979).Article 

    Google Scholar 
    Salt, G. & van den Bosch, R. The defense reactions of three species of Hypera (Coleoptera, curculionidae) to an ichneumon wasp. J. Invertebr. Pathol. 9, 164–177 (1967).Article 

    Google Scholar 
    Yaninek, J. S. & Bellotti, A. C. Exploration for natural enemies of cassava green mites based on agrometeorological criteria. In Proceedings of the Seminar on Agrometeorology and Crop Protection in the Lowland Humid and Subhumid Tropics, Cotonou, Benin, 7–11 July 1986 (eds Rijks, D. & Mathys, G.) 69–75 (World Meteorological Organization, Geneva, Switzerland, 1987).Yaninek, J. S., Onzo, A. & Ojo, J. B. Continent-wide releases of neotropical phytoseiids against the exotic cassava green mite in Africa. Exp Appl Acarol 17, 145–160 (1993).Article 

    Google Scholar 
    Shameer, K. & Hardy, I. C. Host–parasitoid trophic webs in complex agricultural systems. Curr. Opin. Insect Sci. 65, 101253 (2024).Article 
    PubMed 

    Google Scholar 
    Cure, J. R., Rodríguez, D., Gutierrez, A. P. & Ponti, L. The coffee agroecosystem: bio-economic analysis of coffee berry borer control (Hypothenemus hampei). Sci. Rep. 10, 12262 (2020).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Abram, P. K., Mills, N. J. & Beers, E. H. Review: classical biological control of invasive stink bugs with egg parasitoids—What does success look like?. Pest Manag. Sci. 76, 1980–1992 (2020).Article 
    PubMed 

    Google Scholar 
    Gutierrez, A. P. et al. Tritrophic analysis of the prospective biological control of brown marmorated stink bug, Halyomorpha halys, under extant weather and climate change. J. Pest Sci. https://doi.org/10.1007/s10340-023-01610-y (2023).Article 

    Google Scholar 
    Gutierrez, A. P., Ponti, L., Cristofaro, M., Smith, L. & Pitcairn, M. J. Assessing the biological control of yellow starthistle (Centaurea solstitialis L): prospective analysis of the impact of the rosette weevil (Ceratapion basicorne (Illiger)). Agric. For. Entomol. 19, 257–273 (2017).Article 

    Google Scholar 
    Smith, L. USDA Petition to Release Proposed Field Release of the Weevil, Ceratapion basicorne (Coleoptera: Apionidae), from Turkey for Biological Control of Yellow Starthistle (Centaurea solstitialis) in the United States. (mailed 26 January 2006) (2006).Sutton, G. F., Canavan, K., Day, M. D. & Paterson, I. D. Field-based ecological studies to assess prospective biological control agents for invasive alien plants: An example from giant rat’s tail grass. J. Appl. Ecol. 58, 1043–1054 (2021).Article 

    Google Scholar 
    Elith, J. & Leathwick, J. R. Species distribution models: ecological explanation and prediction across space and time. Annu. Rev. Ecol. Evol. Syst. 40, 677–697 (2009).Article 

    Google Scholar 
    Williams, C. L., Goldson, S. L., Baird, D. B. & Bullock, D. W. Geographical origin of an introduced insect pest, Listronotus bonariensis (Kuschel), determined by RAPD analysis. Heredity 72, 412–419 (1994).Article 
    PubMed 

    Google Scholar 
    Barlow, N. D. Models in biological control: a field guide. In Theoretical approaches to biological control (eds Hawkins, B. A. & Cornell, H. V.) 43–70 (Cambridge University Press, 1999).Chapter 

    Google Scholar 
    Gutierrez, A. P. & Ponti, L. Eradication of invasive species: why the biology matters. Environ. Entomol. 42, 395–411 (2013).Article 
    PubMed 

    Google Scholar 
    Gutierrez, A. P., Ponti, L., Herren, H. R., Baumgärtner, J. & Kenmore, P. E. Deconstructing Indian cotton: Weather, yields, and suicides. Environ. Sci. Eur. 27, 12 (2015).Article 

    Google Scholar 
    Simmons, G. et al. Psymple: A python package for complex systems modelling. J. Open Sour. Softw. 10, 7364 (2025).Article 
    ADS 

    Google Scholar 
    Nagel, T. Mind and cosmos: Why the materialist neo-Darwinian conception of nature is almost certainly false (Oxford University Press, 2012).Book 

    Google Scholar 
    Legg, J. P. et al. Cassava virus diseases: Biology, epidemiology, and management. Adv. Virus Res. 91, 85–142 (2015).Article 
    PubMed 

    Google Scholar 
    Mills, N. J. & Gutierrez, A. P. Prospective modelling in biological control: an analysis of the dynamics of heteronomous hyperparasitism in a cotton-whitefly-parasitoid system. J. Appl. Ecol. 33, 1379–1394 (1996).Article 

    Google Scholar 
    Regev, U., Gutierrez, A. P., Schreiber, S. J. & Zilberman, D. Biological and economic foundations of renewable resource exploitation. Ecol. Econ. 26, 227–242 (1998).Article 

    Google Scholar 
    Gutierrez, A. P. The physiological basis of ratio-dependent predator-prey theory: The metabolic pool model as a paradigm. Ecology 73, 1552–1563 (1992).Article 

    Google Scholar 
    Gutierrez, A. P. & Regev, U. The bioeconomics of tritrophic systems: applications to invasive species. Ecol. Econ. 52, 383–396 (2005).Article 

    Google Scholar 
    Heisenberg, W. Physics and philosophy (Harper Torchbooks, 1958).
    Google Scholar 
    Capra, F. & Luisi, P. L. The systems view of life: A unifying vision (Cambridge University Press, 2014).Book 

    Google Scholar 
    Jørgensen, S. E. & Nielsen, S. N. Use of eco-exergy in ecological networks. Ecol. Model. 293, 202–209 (2014).Article 

    Google Scholar 
    Cock, J. H., Franklin, D., Sandoval, G. & Juri, P. The ideal cassava plant for maximum yield. Crop Sci. 19, 271–279 (1979).Article 

    Google Scholar 
    de Wit, A. et al. 25 years of the WOFOST cropping systems model. Agric. Syst. 168, 154–167 (2019).Article 

    Google Scholar 
    Manetsch, T. J. Time-varying distributed delays and their use in aggregative models of large systems. IEEE Trans. Syst. Man Cybern. 6, 547–553 (1976).Article 
    ADS 

    Google Scholar 
    Vansickle, J. Attrition in distributed delay models. IEEE T. Syst. Man Cyb. 7, 635–638 (1977).Article 

    Google Scholar 
    Di Cola, G., Gilioli, G. & Baumgärtner, J. Mathematical models for age-structured population dynamics. In Ecological entomology (eds Huffaker, C. B. & Gutierrez, A. P.) (Wiley, New York, USA, 1999).Buffoni, G. & Pasquali, S. Structured population dynamics: continuous size and discontinuous stage structures. J. Math. Biol. 54, 555–595 (2007).Article 
    MathSciNet 
    PubMed 

    Google Scholar 
    de Wit, C. T. & Goudriaan, J. Simulation of ecological processes Vol. 2nd (Pudoc, 1978).
    Google Scholar 
    Gutierrez, A. P., Falcon, L. A., Loew, W., Leipzig, P. A. & Van Bosch, R. An analysis of cotton production in California: A model for Acala cotton and the effects of defoliators on its yields. Environ. Entomol. 4, 125–136 (1975).Article 

    Google Scholar 
    Wang, Y., Gutierrez, A. P., Oster, G. & Daxl, R. A population model for plant growth and development: Coupling cotton-herbivore interaction. Can. Entomol. 109, 1359–1374 (1977).Article 

    Google Scholar 
    Dreyer, H. & Baumgärtner, J. The influence of post-flowering pests on cowpea seed yield with particular reference to damage by Heteroptera in southern Benin. Agr. Ecosyst. Environ. 53, 137–149 (1995).Article 

    Google Scholar 
    Ruane, A. C., Goldberg, R. & Chryssanthacopoulos, J. Climate forcing datasets for agricultural modeling: Merged products for gap-filling and historical climate series estimation. Agric. For. Meteorol. 200, 233–248 (2015).Article 
    ADS 

    Google Scholar 
    Neteler, M., Bowman, M. H., Landa, M. & Metz, M. GRASS GIS: A multi-purpose Open Source GIS. Environ. Model. Softw. 31, 124–130 (2012).Article 

    Google Scholar 
    GRASS Development Team. Geographic Resources Analysis Support System (GRASS) Software, Version 8.2.0. (Open Source Geospatial Foundation. URL http://grass.osgeo.org, Beaverton, Oregon, USA, 2022). https://doi.org/10.5281/zenodo.5176030.Gutierrez, A. P. & Ponti, L. Pascal code for the cassava tri-trophic PBDM system. Zenodo https://doi.org/10.5281/zenodo.17559583 (2025).Download referencesAcknowledgementsTwo generous reviewers clarified important concepts in biological control that improved the paper. We are grateful to the international network maintaining the Geographic Resources Analysis Support System (GRASS) for making the GIS software available to the scientific community.FundingThe study was supported by CASAS Global NGO (https://www.casasglobal.org), the McKnight Foundation (grants number 22-341 and 24-124), and project TEBAKA (project ID: ARS01_00815) co-funded by the European Union—ERDF and ESF, “PON Ricerca e Innovazione 2014–2020”.Author informationAuthors and AffiliationsCenter for the Analysis of Sustainable Agroecological Systems, Kensington, CA, 94707-1035, USAAndrew Paul Gutierrez & Luigi PontiDivision of Ecosystem Science, College of Natural Resources, University of California, Berkeley, CA, 94720-3114, USAAndrew Paul GutierrezAgenzia nazionale per le nuove tecnologie, l’energia e lo sviluppo economico sostenibile (ENEA), Centro Ricerche Casaccia, Via Anguillarese 301, 00123, Roma, ItalyLuigi PontiInternational Institute of Tropical Agriculture, Biological Control Center for Africa, 08 B.P. 0932, Cotonou, BeninPeter NeuenschwanderDepartment of Entomology, Purdue University, West Lafayette, IN, 47907-2089, USAJohn S. YaninekThe Millennium Institute, Washington, DC, USAHans R. HerrenAuthorsAndrew Paul GutierrezView author publicationsSearch author on:PubMed Google ScholarLuigi PontiView author publicationsSearch author on:PubMed Google ScholarPeter NeuenschwanderView author publicationsSearch author on:PubMed Google ScholarJohn S. YaninekView author publicationsSearch author on:PubMed Google ScholarHans R. HerrenView author publicationsSearch author on:PubMed Google ScholarContributionsAll authors contributed equally.Corresponding authorCorrespondence to
    Andrew Paul Gutierrez.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Ethics approval and consent to participate
    Not applicable.

    Consent for publication
    Not applicable.

    Additional informationPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary InformationSupplementary Information.Rights 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 articleGutierrez, A.P., Ponti, L., Neuenschwander, P. et al. Predicting natural enemy efficacy in biological control using ex-ante analyses.
    Sci Rep 15, 44886 (2025). https://doi.org/10.1038/s41598-025-29022-1Download citationReceived: 25 August 2025Accepted: 13 November 2025Published: 29 December 2025Version of record: 29 December 2025DOI: https://doi.org/10.1038/s41598-025-29022-1Share 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
    KeywordsBiological controlPhysiologically based demographic modelTime-varying life tablesBioeconomic analysis More

  • in

    Succession of micro-food webs and their role in lignocellulose degradation during pepper stalk composting

    AbstractAbtract. Multitrophic interactions among bacteria, fungi, protists, and nematodes play vital roles in organic matter decomposition in soil ecosystems, yet their contributions during composting remain poorly understood. Improving our understanding of these cross-trophic dynamics is essential for optimizing microbial regulation in composting systems. In this study, we investigated the community dynamics and interactions of these trophic groups throughout the aerobic composting of pepper stalks and evaluated their collective influence on lignocellulose degradation. Bacteria expanded rapidly in the early phase, dominating organic matter decomposition, but declined during maturation. Fungi remained low in abundance, with Aspergillus transiently dominating the thermophilic stage. Phagotrophic protists shifted from Colpodella to stress-tolerant Oxytrichidae, while nematodes, which were absent at peak temperature, recovered later, shifting from Rhabditella to Panagrolaimus. Pronounced temporal shifts in community composition and diversity were observed. Co-occurrence network analysis showed increasing interaction complexity over time, with bacterial and fungal taxa dominating lignocellulose-associated modules. Protists and nematodes exerted top-down effects via trophic cascades, indirectly enhancing microbial activity. Structural equation modeling confirmed that food web complexity mediated the indirect effects of community structure on lignocellulose degradation. These findings provide novel insights into compost multitrophic dynamics and highlight the ecological importance of cross-trophic interactions in optimizing decomposition efficiency.

    Similar content being viewed by others

    Microbial consortia-mediated rice residue decompositon for eco-friendly management

    Article
    Open access
    05 September 2025

    Multiyear analysis uncovers coordinated seasonality in stocks and composition of the planktonic food web in the Baltic Sea proper

    Article
    Open access
    22 July 2023

    Microbially mediated mechanisms underlie soil carbon accrual by conservation agriculture under decade-long warming

    Article
    Open access
    08 January 2024

    IntroductionPeppers (Capsicum spp.) are one of the most important crops, with China accounting for approximately one-third of the world’s pepper production1. In 2019, China’s pepper cultivation area reached 21,474 square kilometers, making it the largest vegetable crop by planting area in the country2. The large-scale cultivation of peppers generates a substantial amount of waste, including pepper stalks and other residues. Disposal methods such as discarding or burning in fields not only waste valuable organic material but also contribute to environmental pollution.Aerobic composting, a process of controlled biological decomposition under oxygen-rich conditions, offers a promising solution for managing pepper stalk waste3. This method exposes organic residues to microbial activity, enabling thermophilic decomposition and maturation4. Aerobic composting accelerates organic matter degradation, eliminates pathogens and parasite eggs, and converts organic materials into stable humus, making it a highly efficient and eco-friendly approach to waste management5,6. However, pepper stalks contain a large amount of lignocellulose (composed of lignin, cellulose, and hemicellulose), which is highly resistant to microbial degradation7. This recalcitrance extends the processing time during composting and hinders the production of humus8. To address these limitations, co-composting with other materials such as manure and wheat straw is commonly employed. These amendments provide easily degradable organic matter and low-lignification components, enhancing the decomposition efficiency of pepper stalks9.Bacteria and fungi are the primary drivers of organic matter decomposition during the composting process, and as such, they have garnered significant research interest. Bacteria, the most abundant microorganisms in compost, exhibit rapid growth and the ability to utilize easily degradable organic matter such as starch and proteins10,11. Fungi, on the other hand, are involved in breaking down the more recalcitrant components of organic waste. For example, some fungi within the Ascomycota and Basidiomycota phyla produce cellulase-like enzymes that effectively degrade lignin and cellulose12. Furthermore, fungi contribute to the physical disruption of materials through hyphal penetration, thereby enhancing the breakdown of lignocellulose13. In addition to bacteria and fungi, phagotrophic protists and nematodes, as key compost biota, also play crucial roles in regulating the composting process, yet they have often been overlooked in composting studies. In soil ecosystems, these organisms constitute integral components of the micro-food web, a multitrophic framework that describes the flow of energy and nutrients among microbes and their predators14. Although this concept has been extensively applied in soil ecology to elucidate decomposition dynamics and nutrient cycling15,16, its relevance to composting systems has received far less attention. Composting, however, represents a highly dynamic, microbially driven process analogous to soil systems, where similar trophic interactions occur under accelerated and thermophilic conditions. Integrating the micro-food web perspective into compost ecology may therefore offer a more comprehensive understanding of biotic interactions and their regulatory roles during organic matter decomposition. By preying on bacteria and fungi, phagotrophic protists help regulate microbial communities in compost. However, research on protists in composting remains limited. Liu et al. (2024) identified Ciliophora as key taxa associated with compost maturity during cow manure composting17. Similarly, Yin et al. (2022) reported that phagotrophic protists were critical predators regulating the structure of denitrifying bacterial communities during the co-composting of pig manure and wheat straw18. These studies underscore the significant roles of phagotrophic protists in compost systems. Nematodes, spanning all trophic levels from primary consumers to specialized predators, play a critical role in the micro-food web19. They are classified into functional groups such as bacterial-feeders, fungal-feeders, plant-feeders, and omnivores20. Based on their r and k life-strategy characteristics, nematodes are classified into cp-groups21. Throughout the composting process, significant changes occur in nematode species composition, life strategies, and feeding behaviors. Bacterial-feeding enrichment opportunists (cp-1) dominate the thermophilic phase, followed by bacterial- and fungal-feeding general opportunists (cp-2), while predator and fungal-feeding nematodes become more prominent in the cooling and maturation stages22. These shifts in nematode community structure can have profound effects on microbial populations and their activity23. Together with bacteria, fungi, and phagotrophic protists, nematodes form a dynamic micro-food web that drives organic matter decomposition and nutrient turnover in compost. However, current research has predominantly focused on changes in individual trophic levels during the composting process, with limited attention given to the overall response of the micro-food web and its regulatory role in the composting process.Recent advances in ecological network analysis have provided powerful tools to quantify the complex web of interactions among microorganisms and higher trophic groups24,25. In this framework, microbial or multitrophic networks are constructed based on correlations or co-occurrence patterns among taxa, revealing potential ecological interactions such as competition, mutualism, or predation26. Network complexity, typically expressed through metrics such as the number of nodes and edges and network connectivity, reflects the degree of interdependence and stability within biological communities27,28. In compost ecosystems, where microbial communities experience rapid turnover and strong environmental fluctuations, changes in network complexity can indicate shifts in system stability and functional coordination among decomposer guilds29.To fill this research gap, we focus on the composting of pepper stalks to investigate the dynamic changes of micro-food web attributes (multitrophic diversity, multitrophic composition and network complexity) during aerobic composting, as well as their impact on lignocellulose degradation. We hypothesized that: (i) multitrophic diversity, community composition, and network complexity exhibit significant dynamic changes during the composting process; (ii) the degradation of lignocellulose is regulated by the complexity of micro-food web interaction.Materials and methodsComposting materialsThe experimental site was located at the Key Laboratory of Fertilization from Agricultural Wastes, Hubei Academy of Agricultural Sciences, Wuhan. The composting materials included pepper straw, pig manure and wheat straw, with pig manure providing essential nitrogen and microbial inoculum, while wheat straw contributed to adjusting the C/N ratio and improving porosity. The pepper straw was sourced from the Industrial Crops Institute, Hubei Academy of Agricultural Sciences; the pig manure was obtained from the Jinshui Base of the Hubei Academy of Agricultural Sciences; the wheat straw was provided by Hubei Guangmei Technology Co., Ltd. The physicochemical properties of these materials are shown in Table S1 (Supplementary Material).Composting experiments and sample collectionThe composting system employed was a customized VTD-100 reactor manufactured by Qinhuangdao Nuoxin Environmental Protection Technology Co., Ltd. (China). Constructed from stainless steel, the reactor was insulated with a 3 cm-thick thermal layer and had an effective capacity of 100 L (Fig. S1, Supplementary Material). The experiment was set up with four independent replicates. The pepper straw, pig manure, and wheat straw were mixed at a ratio of 7:1:1.25 (w/w), with the carbon to nitrogen (C/N) ratio adjusted to 25 and the moisture content adjusted to 55%. According to the temperature profile, the composting process proceeded for a total of 26 days. Turning was conducted on composting days 4, 6, 8, 9, 13, and 20 to enhance oxygen content in the materials. The temperature of the core fermentation zone of the composting pile was measured at 9:00 and 16:00 daily using a stainless-steel compost thermometer, and the average value was taken as the daily pile temperature, while also recording the ambient temperature.Compost samples were collected at five representative time points (Days 0, 2, 7, 10, and 26) corresponding to distinct thermal phases (initial, mesophilic, thermophilic, cooling, and maturation). Samples were obtained from five positions (four corners and the center) at approximately 20 cm depth, targeting the core fermentation zone where microbial activity and temperature were most stable. Subsamples were pooled and homogenized to form a composite sample (~ 1000 g). All samples were divided into three subgroups: one portion was air-dried for physicochemical analyses; another portion was used fresh to determine moisture content and for nematode extraction; and a third portion was used for microbial analysis (stored at −80 °C).Analysis of composting physicochemical propertiesTotal organic carbon (TOC) was determined using the potassium dichromate external heat source method. H2SO4-H2O2 digestion was used for the determination of total nitrogen (TN), total phosphorus (TP) and total potassium (TK). TN was determined by the semi-micro Kjeldahl method, TP by the vanadium-molybdate yellow colorimetric method, and TK by flame photometry. The pH and electrical conductivity (EC) were determined using a 1:10 solid-liquid extraction ratio; The pH was measured with a pH meter (INESA, PHS-3E, China), and EC was measured with a conductivity meter (OHAUS, ST3100C, China). Ash content was measured using the muffle furnace incineration method (550 °C for 6 h). Lignocellulose content was determined by Van Soest detergent fiber analysis method, using the automatic cellulose analyzer (Fibertec™ 8000, Denmark) to measure neutral detergent fiber (NDF), acid detergent fiber (ADF), and acid detergent lignin (ADL). Hemicellulose content was calculated as the difference between NDF and ADF, cellulose content as the difference between ADF and ADL, and lignin content as the difference between ADL and ash content. The degradation rate of lignocellulose was calculated as follows:Degradation rate (%) = (C0 − Ct)/C0 × 100.Where C0 is the initial content and Ct represents the residual content at sampling time point.DNA extraction, quantitative PCR and high-throughput sequencingDNA was extracted using the MoBio PowerSoil DNA Isolation kit, and the concentration and quality of DNA were measured by the NanoDrop 2000 (Thermo Fisher Scientific). The abundance of bacteria and fungi was quantified using quantitative PCR (qPCR) on an ABI Prism 7500 cycler (Applied Biosystems, Germany), with primers 338 F/806R30, and ITS1F/ITS2R31, respectively. Reaction mixtures (20 µl) contained 10 µL of 2×ChamQ SYBR Color qPCR Master Mix, 0.4 µl 50×ROX Reference Dye 1, 0.8 µl of each primer (5 µM), 2 µl of template DNA (10 ng/µl) and 6 µl of ddH2O. PCR conditions were 3 min at 95 °C, followed by 40 cycles of 95 °C for 5 s, 58 °C for 30 s, and 72 °C for 1 min. Each plate included triplicate reactions per DNA sample, the appropriate set of standards and negative controls.The V3-V4 region of bacterial 16S rRNA gene, ITS2 region of fungal rRNA operon and V4 region of protist 18S rRNA gene were amplified using the primers 338F/806R (5’-ACTCCTACGGGAGGCAGCAG-3’, 5’-GGACTACHVGGGTWTCTAAT-3’), ITS1F/ITS2R (5’-CTTGGTCATTTAGAGGAAGTAA-3’, 5’-GCTGCGTTCTTCATCGATGC-3’), and TAReuk454F/TAReukR (5’-CCAGCASCYGCGGTAATTCC-3’, 5’-ACTTTCGTTCTTGATYRA-3’)32, respectively. The high-throughput sequencing of bacteria, fungi and phagotrophic protists were conducted on the Illumina MiSeq PE300 platform at Shanghai Majorbio Biotechnology Co. Ltd, (Shanghai, China). The raw sequences were processed using the QIIME1 pipeline. Briefly, the quality control of the raw sequences was examined by fastp33, and pair-end sequences were merged using FLASH34. Operational taxonomic units (OTUs) of bacterial, fungal and phagotrophic protists sequences were identified using UCLUST at a 97% similarity cutoff. Taxonomic classification was conducted based on the SILVA v138 database for bacteria, the UNITE database 8.0 for fungi and the protist ribosomal reference database 4.5 for protist35. The protist OTU table was generated by removing OTUs classified as Rhodophyta, Streptophyta, Metazoa, Fungi, Opisthokonta_X, and unclassified taxa35. Protist functional groups were manually classified into phagotrophic and phototrophic categories based on their feeding modes. In this study, only phagotrophic protists were considered. Sequences were rarefied to the minimum sequencing depth at 26, 671 for bacteria, 32, 807 for fungi and 157 for phagotrophic protist.Nematode community analysisNematodes were extracted from 50 g of fresh compost using the Baermann funnel method36. Briefly, each sample was placed on a double layer of tissue paper inside a glass funnel filled with water, ensuring that the sample was fully submerged without direct contact with the bottom. The setup was left undisturbed at room temperature (approximately 20–25 °C) for 24–48 h to allow nematodes to migrate into the water and settle at the bottom of the stem. Afterward, the nematode suspension was collected from the bottom of the funnel, and the total number was counted under a microscope. Then, 100 individuals (or all if fewer than 100) were randomly selected and identified to the genus level using an optical microscope.Statistics analysisShannon index was calculated using the “vegan” package in R (version 4.0.2) to characterize the α-diversity of microbial communities. To represent multitrophic diversity, the shannon diversity indices of bacteria, fungi, phagotrophic protists and nematodes were first standardized (Z-score transformation) and then averaged37. Multitrophic composition was visualized through principal coordinates analysis (PCoA) based on genus-level Bray-Curtis dissimilarity matrices of bacterial, fungal, phagotrophic protist, and nematode communities using the “vegan” package, and the first axis of principal coordinates analysis (PCoA1) was used to quantify the multitrophic composition. The physicochemical parameters and α-diversity were analyzed using one-way analysis of variance (ANOVA) to evaluate differences across composting stages, with significant differences identified through Student-Newman-Keuls test (P < 0.05) using the SPSS v20.0. Co-occurrence networks were constructed based on Spearman correlation coefficients (r > 0.8, P < 0.01) to assess the interaction patterns among bacteria, fungi, phagotrophic protists and nematodes, and visualized by Gephi 0.10.1. The distribution patterns of the major modules were evaluated by standardizing and averaging the relative abundance of OTUs in each module38. Topological parameters (include node number, edge number and average weighted degree) of sub-networks were calculated to evaluate network complexity and its temporal dynamics during composting using the subgraph function of the “igraph” package in R39. Linear regression was employed to analyze the relationships between multitrophic diversity, composition, and complexity (average weighted degree) of micro-food web with degradation rates of lignocellulose using “ggplot2” package in R. The structural equation model (SEM) was constructed to evaluate the direct and indirect effects of composting pile properties (temperature, pH) on lignocellulose degradation through multitrophic diversity, multitrophic composition and micro-food web complexity. The model fit of the SEM was evaluated by a chi-square test (P > 0.05), the goodness-of-fit index (GFI > 0.9), and the root mean square error of approximation (RMSEA < 0.05).Results and discussionPhysicochemical process and lignocellulose degradation during compostingA typical composting process consists of mesophilic, thermophilic (> 45 °C), cooling and maturation phase40. In our study, the compost pile entered the thermophilic phase on Day 1, reached a maximum temperature of 64.1 °C on Day 2, and remained above 45.0 °C for 5 days (Fig. 1A). This sharp rise in temperature was attributed to the rapid biodegradation of easily degradable organic matter, which provided an abundant energy source for microbial metabolism during the initial phase of composting41. As the supply of organic matter diminished, microbial activity and heat generation gradually decreased, resulting in a steady decline in temperature on Day 6. By Day 26, the temperature had stabilized, approaching that of the ambient temperature, indicating the compost had reached the maturation phase. During the composting process, the pH value continuously increased due to ammonification42, reaching 9.6 by day 10, after which it remained stable (Fig. S2, Supplementary Material). The EC value initially declined but began rising after Day 2, continuing to increase due to the degradation of organic matter, which concentrated soluble salts43. In the early stages of composting (Day 2 to Day 7), TOC content significantly decreased due to the rapid decomposition of easily degradable organic matter by microorganisms. Afterward, the rate of TOC reduction slowed down. Meanwhile, TN, TP and total TK content levels increased rapidly between day 2 and day 10, followed by a slower rise. After composting, TN, TP, and TK contents increased from 1.38%, 0.71%, and 2.43% to 2.00%, 0.95%, and 3.18%, respectively. By day 26, the C/N ratio decreased to 20.8.As shown in Fig. 1B, the degradation rate of lignocellulosic components increased progressively throughout the composting process. During the thermophilic phase, the degradation rate of lignin and hemicellulose increased rapidly, while cellulose degradation remained relatively low. Notably, between Day 7 and Day 10, the degradation amount of cellulose exhibited a significant increase. In contrast, lignin degradation slowed considerably after Day 7. By the end of the composting, the degradation rate for cellulose, hemicellulose, and lignin were 5.75%, 7.47%, and 3.74%, respectively. Lignocellulose, being a high-molecular-weight organic compound, is relatively difficult for microorganisms to degrade8,44. During the composting process, readily degradable organic carbon sources such as starch, proteins, and sugars are preferentially metabolized by microorganisms45. As a result, the degradation rate of lignocellulose is initially low. The random, amorphous structure of hemicellulose makes it highly susceptible to microbial degradation, making it the most easily degradable substrate46,47. The slow degradation of cellulose in the early stages occurs because it is encapsulated by lignin, while hemicellulose forms a structural connection between lignin and cellulose fibers, further hindering cellulase binding44. From Day 7 to Day 10, the accelerated degradation of hemicellulose facilitated the degradation of cellulose. Compared to cellulose and hemicellulose, lignin has a more complex structure as an amorphous aromatic compound with a three-dimensional configuration, making it more resistant to microbial degradation48. Ligninolytic enzymes such as laccase and peroxidases are crucial for lignin degradation, and their activities are enhanced under thermophilic conditions49,50. After Day 7, the compost temperature decreased, resulting in a significantly reduced amount of lignin degradation.Fig. 1Temperature dynamics (A) and lignocellulose degradation (B) during the composting process.Full size imageSuccession of bacterial and fungal communities during compostingBacterial abundance increased rapidly in the early composting phase, indicating a dominant role in organic matter decomposition, but declined sharply by Day 26 (Fig. 2A). This pattern is consistent with previous findings that bacterial populations expand during active decomposition and contract during maturation due to substrate depletion51. Bacterial diversity was initially low on Day 0 but gradually increased throughout composting (Fig. 2B). At the genus level (Fig. 2C), Pantoea dominated the initial community (24.3%), while Bacillus became predominant during the thermophilic phase (21.8% on Day 2). Oceanobacillus also peaked at this stage (12.8%). The Bacillus genus is highly adaptable, with species that reproduce rapidly and can form spores under harsh conditions. Their ability to degrade complex organic matter, such as cellulose, and their heat tolerance make them effective functional inoculants in composting52,53. Similarly, due to its tolerance to high temperature and salinity stress, along with its nitrifying capacity, Oceanobacillus is well-suited for organic composting54. During the cooling phase, Saccharomonospora increased on Day 7 (15.9%) but declined by Day 26 (6.8%). As a thermotolerant Actinobacteria genus, it contributes to compost maturation by converting lignin and cellulose into humic substances and hydrolyzing phenolic compounds into less toxic forms, thereby reducing phytotoxicity and improving compost quality55,56.In contrast, fungal abundance was much lower than bacterial abundance, rapidly declining after Day 0 and remaining low throughout composting (Fig. 2D). This decline might be attributed to the faster metabolism of bacteria, which makes them more competitive than fungi. Additionally, bacterial higher surface-to-volume ratio, greater diversity, and shorter generation times enable them to better adapt to the swift fluctuations in substrate and environmental conditions throughout the composting process57. Fungal diversity remained relatively stable throughout the composting process (Fig. 2E), but notable shifts in community composition were observed (Fig. 2F). At the genus level, Aspergillus was predominant and exhibited marked temporal dynamics. It accounted for only 5.7% on Day 0 but rapidly increased to dominate the community between Day 2 and Day 10, with relative abundance ranging from 47.6% to 66.6%. By the end of composting, its abundance declined to 15.3%, while Hormographiella became dominant (42.5%). The transient dominance of Aspergillus during the thermophilic phase is consistent with its known role in lignocellulose degradation through the production of multi-enzymatic complexes capable of efficiently breaking down plant biomass58.Fig. 2The dynamic of bacterial and fungal communities during the composting process. (A) Quantification of bacterial abundance; (B) Changes in bacterial diversity (shannon index) during composting; (C) Relative abundance of dominant bacterial genus; (D) Quantification of fungal abundance; (E) Changes in fungal diversity (shannon index) during composting; (F) Relative abundance of dominant fungal genus. Different letters indicate significant differences (P < 0.05).Full size imageSuccession of higher trophic organisms and the multitrophic community during compostingHigher trophic organisms also exhibited pronounced temporal changes during composting. Phagotrophic protist diversity was high on Days 0 and 2 but declined markedly by Days 10 and 26 (Fig. 3A). Early in the process (Days 0–2), the community was primarily dominated by Colpodella (Fig. 3B), a genus of free-living, predatory flagellates that shares synapomorphic features with parasitic Apicomplexa59. As composting progressed, an unidentified genus from the family Oxytrichidae gradually increased in abundance and became dominant by Days 10 and 26. Members of Oxytrichidae, a family of bacterivorous ciliates in the order Sporadotrichida, are capable of forming resistant cysts, enabling survival under dry or otherwise unfavorable conditions60.Nematodes were also detected throughout the composting process. Although ubiquitous in compost ecosystems22,61, their abundance was highly temperature-dependent. Consistent with prior studies indicating a thermal threshold around 40 °C for nematode activity19, we observed a complete absence of nematodes at the peak thermophilic stage (Day 2), followed by a gradual recovery as temperatures declined, peaking at Day 26 (Fig. 3C). This pattern suggests that nematodes may have survived the thermophilic peak in the form of dormant stages or eggs, or found refuge in cooler outer zones of the compost pile22. Only two genera, Rhabditella and Panagrolaimus, were detected using the shallow plate method (Fig. 3D), likely due to the biofumigation effect of Capsicum annuum residues, which are known to suppress plant-parasitic nematodes62. Both Rhabditella and Panagrolaimus are bacterial-feeding enrichment opportunists (cp-1), well-adapted to early compost environments19. Rhabditella was dominant at Day 0, while Panagrolaimus became predominant from Day 7 onward.Fig. 3The dynamic of phagotrophic protist and nematode community during the composting process. (A) Changes in phagotrophic protist diversity (shannon index) during composting; (B) Relative abundance of dominant phagotrophic protist genus; (C) Changes in nematode abundance (individuals per 100 g dry compost) during composting; (D) Relative abundance of nematode genus. Different letters indicate significant differences (P < 0.05).Full size imageThe successional shifts in both phagotrophic protists and nematodes suggest dynamic changes in predator–prey interactions and energy transfer within the compost micro-food web. When integrating all trophic groups, significant temporal shifts in overall multitrophic diversity and community composition were observed. As shown in Fig. 4A, multitrophic diversity was lowest at Day 0, increasing sharply by Day 2 and then remaining relatively stable through Day 26. PCoA revealed clear compositional divergence across time points, with distinct clusters corresponding to Days 0, 2, 7, 10, and 26 (Fig. 4B). Axis 1 explained 51.91% of the variance, and Axis 2 explained 24.77%, indicating that composting had a strong effect on community succession across trophic levels. These temporal shifts in multitrophic diversity and composition reflect not only species turnover but also changes in ecological interactions such as competition, predation, and facilitation. The rapid increase in diversity by Day 2 suggests a swift community reassembly following initial disturbance, likely driven by temperature, substrate availability, and oxygen gradients63. The stabilization of diversity from Day 2 onward may indicate the formation of a relatively resilient and functionally integrated micro-food web. Moreover, the clear separation of communities across time points in PCoA underscores the stage-specific structuring of trophic interactions, highlighting the importance of temporal niche differentiation in shaping food web complexity during composting64.Fig. 4The dynamic of multitrophic diversity (A) and composition (B) during the composting process.Full size imageStructural development of micro-food web complexity during compostingNetwork analysis has emerged as a powerful tool for unraveling complex and dynamic interactions among microbial taxa in environmental systems26,65. To characterize the temporal dynamics of trophic interactions within the compost micro-food web, a multi-kingdom co-occurrence network was constructed based on bacteria, fungi, phagotrophic protists, and nematodes (Fig. 5A). The results revealed that bacteria dominated the network in terms of node abundance, underscoring their central role in organic matter decomposition. Despite the detection of only two nematode genera in the compost, Panagrolaimus remained within the filtered network (Spearman r > 0.8, P < 0.01) and was significantly associated with 38 nodes, including 33 bacterial taxa. This suggests that, even with limited diversity, nematodes can exert notable ecological influence by mediating trophic interactions and modulating microbial activity66.Modules—clusters of tightly connected nodes—can represent ecological niches or functional microbial units67. Three major modules (Modules 1–3) were identified in the multi-trophic network (Fig. 5A). Module 1 was the largest and most complex, containing 138 nodes and representing a densely interconnected multi-trophic structure (Table S2, Supplementary Material). It was dominated by Proteobacteria, Firmicutes, Bacteroidota, and Actinobacteriota, along with phagotrophic protists and Panagrolaimus. The normalized abundance of Module 1 decreased during the thermophilic phase (Day 2) but recovered and peaked by Day 10 (Fig. 5B), suggesting that these taxa are sensitive to extreme heat but play critical roles in restoring ecosystem stability during the cooling phase. Module 2, characterized by high taxonomic diversity and trophic interactions, included Proteobacteria, Bacteroidota, Actinobacteriota, Ascomycota, and several phagotrophic protists (Cercozoa and Ciliophora) (Table S2, Supplementary Material). Its normalized abundance was low in early composting stages but increased significantly after Day 10, peaking at Day 26 (Fig. 5B), indicating its key role in driving organic matter degradation during the mid-to-late phases of composting. Module 3 mainly consisted of Firmicutes and Ascomycota, with consistently positive associations among its members (Table S2, Supplementary Material). This module likely represents cooperative interactions between bacterial and fungal taxa known to initiate the breakdown of simple organic substrates40. The peak abundance of Module 3 on Day 2 emphasizes its importance in the rapid degradation of labile materials during the thermophilic stage. However, its gradual decline over time reflects a reduced contribution as the compost matures and substrates become more recalcitrant.Subnetwork topological analysis revealed pronounced temporal changes in food web complexity68. Subnetwork topological parameters, such as the number of nodes, number of edges, and average degree, steadily increased and peaked during the maturation phase (Fig. 5C), indicating intensified biotic interactions among bacteria, fungi, protists, and nematodes. This increase likely resulted from a drop in compost temperature, which enabled the proliferation of low-trophic-level taxa (e.g., bacteria) and the re-emergence of high-trophic-level organisms such as nematodes. These shifts enhanced both within-group interactions (e.g., bacteria-bacteria) and cross-trophic linkages (e.g., bacteria-protist or bacteria-nematode), contributing to a more integrated and functionally complex food web toward the end of composting.Fig. 5Co-occurrence network of the micro-food web during the composting. (A) Co-occurrence network analysis of the micro-food web. Left: Nodes represent different taxa, color-coded by group (bacteria, fungi, protists, and nematodes). Right: Nodes are grouped by modules (Module 1, Module 2, Module 3, and others), with edges representing positive (red) and negative (green) interactions between taxa. Node size reflects the degree of connectivity. Changes in the normalized averaged abundance of three modules during composting (B). Changes in network topological parameters during composting (C).Full size imageRelationship between micro-food web community attributes and lignocellulose degradationOur results revealed a significant relationship between micro-food web attributes and lignocellulose degradation dynamics during composting. As shown in Fig. 6A, the multitrophic composition showed a significant negative correlation with the degradation rates of hemicellulose (R2 = 0.69, P = 0.000), cellulose (R2 = 0.61, P = 0.000), and lignin (R2 = 0.65, P = 0.000). In contrast, the micro-food web complexity exhibited strong positive correlations with the degradation rates of hemicellulose (R2 = 0.69, P = 0.000), cellulose (R2 = 0.59, P = 0.000), and lignin (R2 = 0.67, P = 0.000). These findings suggest that interactions within the micro-food web facilitate the degradation of lignocellulose. In comparison, multitrophic diversity showed no significant correlations with lignocellulose degradation. This weak association may reflect a decoupling between diversity and functional efficiency under strong environmental filtering, where only a subset of heat-tolerant taxa remain metabolically active during the thermophilic phase69,70,71. During this phase, extreme temperature and limited oxygen further constrain the interconnectivity of micro-food webs. In contrast, as physicochemical conditions become more stable during the cooling and maturation stages, trophic interactions among bacteria, fungi, protists, and nematodes strengthen and play a more pronounced role in regulating substrate decomposition19. These intensified cross-kingdom interactions likely enhance substrate accessibility and enzymatic efficiency through selective grazing and stimulation of microbial turnover.The structural equation model (SEM) (Fig. 6B) further underscored the critical role of interaction complexity in lignocellulose degradation. It showed that compost temperature directly influenced both multitrophic diversity and interaction complexity, while pH exerted a significant impact on community diversity, composition, and network complexity. Critically, neither multitrophic diversity nor composition directly affected lignocellulose degradation—instead, their influence was mediated through micro-food web complexity. This pathway underscores the regulatory importance of ecological interactions in driving decomposition processes. Previous studies have demonstrated that microbial co-occurrence networks—particularly involving bacterial and fungal communities—can shape decomposition rates by modulating functional potential and substrate targeting40. Our findings expand upon this by incorporating higher trophic levels (e.g., protists and nematodes), suggesting that food web complexity at the ecosystem scale is a more comprehensive and sensitive predictor of lignocellulose degradation efficiency. In the compost environment, cross-kingdom interactions (e.g., bacteria-nematodes) may promote substrate accessibility through selective grazing, enzymatic stimulation, and spatial structuring of microbial activity72. Therefore, fostering a complex and well-connected micro-food web could be an effective strategy to accelerate compost maturation and improve lignocellulose decomposition. These findings also have practical implications for composting management. Enhancing the diversity and activity of microfaunal communities through moisture control, proper aeration, or targeted inoculation may improve decomposition efficiency and reduce composting time.Fig. 6Relationship between micro-food web community attributes and lignocellulose degradation. (A) Linear regression of multitrophic diversity, multitrophic composition, and complexity of micro-food web with degradation rates of lignocellulose; (B) Structural equation model (SEM) of key driving factors. The red line indicates a positive correlation, the blue line indicates a negative correlation, and the gray dashed arrows indicate non-significant relationships. Numbers next to the arrows indicate standardized path coefficients. Significance levels: * P < 0.05, ** P < 0.01, and *** P < 0.001.Full size imageConclusionThis study investigated the dynamic changes of the micro-food web during pepper straw composting and its impact on lignocellulose degradation. The results revealed distinct temporal patterns in the abundance, diversity, and species composition of bacterial, fungal, phagotrophic protist, and nematode communities, reflecting the unique contributions of different trophic groups to the composting process. Bacteria and fungi were identified as key drivers of organic matter decomposition, while protists and nematodes indirectly influenced lignocellulose degradation by regulating microbial communities through predation. The significant association between micro-food web interactions and lignocellulose content underscores the critical role of cross-trophic interactions in facilitating lignocellulose breakdown. Notably, the regulatory effects of the micro-food web varied across composting stages, with stronger trophic interactions emerging during the cooling and maturation phases. These findings highlight the importance of multitrophic interactions within the micro-food web during compost maturation and suggest the potential for optimizing composting strategies by modulating these interactions. However, further research is needed to deepen our understanding of the specific functional roles and mechanisms within the micro-food web.

    Data availability

    The datasets generated and/or analyzed during the current study are available in the NCBI SRA database (www.ncbi.nlm.nih.gov/sra) under accession numbers PRJNA1236533 for bacteria, PRJNA1236543 for fungi, and PRJNA1236582 for protist.
    ReferencesZhang, X. M. Genetic diversity of pepper (Capsicum spp.) germplasm resources in China reflects selection for cultivar types and spatial distribution. J. Integr. Agr. 15, 1991–2001 (2016).Article 

    Google Scholar 
    Zou, Z. & Zou, X. Geographical and ecological differences in pepper cultivation and consumption in China. Front. Nutr. 8, 718517 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Zhao, K. et al. Optimizing the management of aerobic composting for antibiotic resistance genes elimination: a review of future strategy for livestock manure resource utilization. J. Environ. Manage. 370, 122766 (2024).Article 
    PubMed 

    Google Scholar 
    Finore, I. et al. Thermophilic bacteria and their thermozymes in composting processes: a review. Chem. Biol. Technol. Agric. 10, 7 (2023).Article 

    Google Scholar 
    Atif, K. et al. Pathogens evolution during the composting of the household waste mixture enriched with phosphate residues and Olive oil mill wastewater. Waste Biomass Valori. 11, 1789–1797 (2018).Article 

    Google Scholar 
    Gao, X. et al. Diversity in the mechanisms of Humin formation during composting with different materials. Environ. Sci. Technol. 53, 3653–3662 (2019).Article 
    ADS 
    PubMed 

    Google Scholar 
    Serrano, L., Marín, F., Gonzalo, A. & Labidi, J. Integral use of pepper stems. Ind. Crops Prod. 40, 110–115 (2012).Article 

    Google Scholar 
    Jurado, M. M. et al. Enhanced turnover of organic matter fractions by microbial stimulation during lignocellulosic waste composting. Bioresour Technol. 186, 15–24 (2015).Article 
    PubMed 

    Google Scholar 
    Ansari, S. A. et al. Additive facilitated co-composting of lignocellulosic biomass waste: approach towards minimizing greenhouse gas emissions—an up-to-date review. Environ. Res. 224, 115529 (2023).Article 
    PubMed 

    Google Scholar 
    Awasthi, M. K., Duan, Y., Liu, T., Awasthi, S. K. & Zhang, Z. Relevance of Biochar to influence the bacterial succession during pig manure composting. Bioresour Technol. 304, 122962 (2020).Article 
    PubMed 

    Google Scholar 
    Wu, D. et al. Reconstruction of core microbes based on producing lignocellulolytic enzymes caused by bacterial inoculation during rice straw composting. Bioresour Technol. 315, 123849 (2020).Article 
    PubMed 

    Google Scholar 
    Wang, S. P. et al. Biochar addition reduces nitrogen loss and accelerates composting process by affecting the core microbial community during distilled grain waste composting. Bioresour Technol. 337, 125492 (2021).Article 
    PubMed 

    Google Scholar 
    Tian, X. et al. Fungal community and cellulose-degrading genes in the composting process of Chinese medicinal herbal residues. Bioresour Technol. 241, 374–383 (2017).Article 
    PubMed 

    Google Scholar 
    De Ruiter, P. C., Neutel, A. M. & Moore, J. C. Energetics, patterns of interaction strengths, and stability in real ecosystems. Science 269, 1257–1260 (1995).Article 
    ADS 
    PubMed 

    Google Scholar 
    Liu, J. et al. Variations in the soil micro-food web structure and its relationship with soil C and N mineralization during secondary succession of subalpine forests. Sci. Total Environ. 879, 163257 (2023).Article 
    ADS 
    PubMed 

    Google Scholar 
    Du Preez, G. et al. Nematode-based indices in soil ecology: application, utility, and future directions. Soil. Biol. Biochem. 169, 108640 (2022).Article 

    Google Scholar 
    Liu, X. et al. Biodiversity and core microbiota of key-stone ecological clusters regulate compost maturity during cow-dung-driven composting. Environ. Res. 245, 118034 (2024).Article 
    PubMed 

    Google Scholar 
    Yin, Y. et al. Phagotrophic protists can change microbial nitrogen conversion patterns during swine manure composting. Biomass Convers. Bior. 14, 517–524 (2022).Article 

    Google Scholar 
    Steel, H. et al. Factors influencing the nematode community during composting and nematode-based criteria for compost maturity. Ecol. Indic. 85, 409–421 (2018).Article 

    Google Scholar 
    Dietrich, P., Cesarz, S., Liu, T., Roscher, C. & Eisenhauer, N. Effects of plant species diversity on nematode community composition and diversity in a long-term biodiversity experiment. Oecologia 197, 297–311 (2021).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bongers, T. The maturity index, the evolution of nematode life history traits, adaptive radiation and cp-scaling. Plant. Soil. 212, 13–22 (1999).Article 

    Google Scholar 
    Steel, H. et al. Survival and colonization of nematodes in a composting process. Invertebr Biol. 132, 108–119 (2013).Article 

    Google Scholar 
    Zhou, J. et al. Effects of bacterial-feeding nematodes on soil microbial activity and the microbial community in oil-contaminated soil. J. Environ. Manage. 234, 424–430 (2019).Article 
    PubMed 

    Google Scholar 
    Ding, L. J., Ren, X. Y., Zhou, Z. Z., Zhu, D. & Zhu, Y. G. Forest-to-cropland conversion reshapes microbial hierarchical interactions and degrades ecosystem multifunctionality at a National scale. Environ. Sci. Technol. 58, 11027–11040 (2024).Article 
    ADS 
    PubMed 

    Google Scholar 
    Gao, S. H. et al. Microplastics trigger soil dissolved organic carbon and nutrient turnover by strengthening microbial network connectivity and cross-trophic interactions. Environ. Sci. Technol. 59, 5596–5606 (2025).Article 
    ADS 
    PubMed 

    Google Scholar 
    Faust, K. & Raes, J. Microbial interactions: from networks to models. Nat. Rev. Microbiol. 10, 538–550 (2012).Article 
    PubMed 

    Google Scholar 
    Chen, W. et al. Soil microbial network complexity predicts ecosystem function along elevation gradients on the Tibetan plateau. Soil. Biol. Biochem. 172, 108766 (2022).Article 

    Google Scholar 
    Cornell, C. R. et al. Land use conversion increases network complexity and stability of soil microbial communities in a temperate grassland. ISME J. 17, 2210–2220 (2023).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Duan, Y. et al. Bacterial community dynamics and co-occurrence network patterns during different stages of biochar-driven composting. Bioresour Technol. 384, 129358 (2023).Article 
    PubMed 

    Google Scholar 
    Dennis, K. L. et al. Adenomatous polyps are driven by microbe-instigated focal inflammation and are controlled by IL-10-producing T cells. Cancer Res. 73, 5905–5913 (2013).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Ji, C. et al. Combined application effects of arbuscular mycorrhizal fungi and Biochar on the rhizosphere fungal community of Allium fistulosum L. J. Microbiol. Biotechnol. 33, 1013–1022 (2023).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Stoeck, T. et al. Multiple marker parallel Tag environmental DNA sequencing reveals a highly complex eukaryotic community in marine anoxic water. Mol. Ecol. 19, 21–31 (2010).Article 
    PubMed 

    Google Scholar 
    Chen, S., Zhou, Y., Chen, Y. & Gu, J. Fastp: an ultra-fast all-in-one FASTQ preprocessor. Bioinformatics 34, i884–i890 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Magoc, T. & Salzberg, S. L. FLASH: fast length adjustment of short reads to improve genome assemblies. Bioinformatics 27, 2957–2963 (2011).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Xu, M. et al. Deciphering the role of biotic factors in community assembly and diversity patterns of soil bacterial communities under different land use. Ecol. Indic. 157, 111273 (2023).Article 

    Google Scholar 
    Hooper, D. J. Extraction of free-living stages from soil. In Laboratory Methods for Work with Plant and Soil Nematodes (ed. Southey, J. F.) 5–30 (Her Majesty’s Stationary Office, 1986).
    Google Scholar 
    Wang, D. et al. Soil micro-food web complexity drives soil multifunctionality along an elevation gradient. Catena 246, 108464 (2024).Article 

    Google Scholar 
    Chun, S. J. et al. Network analysis reveals succession of Microcystis genotypes accompanying distinctive microbial modules with recurrent patterns. Water Res. 170, 115326 (2020).Article 
    PubMed 

    Google Scholar 
    Ma, B. et al. Geographic patterns of co-occurrence network topological features for soil microbiota at continental scale in Eastern China. ISME J. 10, 1891–1901 (2016).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Meng, L., Xu, C., Wu, F. & Huhe Microbial co-occurrence networks driven by low-abundance microbial taxa during composting dominate lignocellulose degradation. Sci. Total Environ. 845, 157197 (2022).Article 
    PubMed 

    Google Scholar 
    Wei, H., Wang, L., Hassan, M. & Xie, B. Succession of the functional microbial communities and the metabolic functions in maize straw composting process. Bioresour Technol. 256, 333–341 (2018).Article 
    PubMed 

    Google Scholar 
    Gao, M. et al. The effect of aeration rate on forced-aeration composting of chicken manure and sawdust. Bioresour Technol. 101, 1899–1903 (2010).Article 
    ADS 
    PubMed 

    Google Scholar 
    Gomez-Brandon, M., Lazcano, C. & Dominguez, J. The evaluation of stability and maturity during the composting of cattle manure. Chemosphere 70, 436–444 (2008).Article 
    ADS 
    PubMed 

    Google Scholar 
    Ma, C. et al. Molecular mechanisms underlying lignocellulose degradation and antibiotic resistance genes removal revealed via metagenomics analysis during different agricultural wastes composting. Bioresour Technol. 314, 123731 (2020).Article 
    PubMed 

    Google Scholar 
    Sun, C. et al. Improve spent mushroom substrate decomposition, bacterial community and mature compost quality by adding cellulase during composting. J. Clean. Prod. 299, 126928 (2021).Article 

    Google Scholar 
    Hendriks, A. T. & Zeeman, G. Pretreatments to enhance the digestibility of lignocellulosic biomass. Bioresour Technol. 100, 10–18 (2009).Article 
    PubMed 

    Google Scholar 
    Serramia, N. et al. Contribution of the lignocellulosic fraction of two-phase olive-mill wastes to the degradation and humification of the organic matter during composting. Waste Manag. 30, 1939–1947 (2010).Article 
    PubMed 

    Google Scholar 
    Harindintwali, J. D., Zhou, J. & Yu, X. Lignocellulosic crop residue composting by cellulolytic nitrogen-fixing bacteria: a novel tool for environmental sustainability. Sci. Total Environ. 715, 136912 (2020).Article 
    PubMed 

    Google Scholar 
    Guo, Y. X. et al. Succession of the microbial communities and function prediction during short-term Peach sawdust-based composting. Bioresour Technol. 332, 125079 (2021).Article 
    PubMed 

    Google Scholar 
    Li, J. et al. Increased temperature enhances microbial-mediated lignin decomposition in river sediment. Microbiome 13, 89 (2025).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Galitskaya, P. et al. Fungal and bacterial successions in the process of co-composting of organic wastes as revealed by 454 pyrosequencing. PLoS One. 12, e0186051 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Zhang, S. et al. Role of Bacillus inoculation in rice straw composting and bacterial community stability after inoculation: unite resistance or individual collapse. Bioresour Technol. 337, 125464 (2021).Article 
    PubMed 

    Google Scholar 
    Wang, L. et al. Enhanced lignocellulose degradation and compost fertility of cattle manure and wheat straw composting by Bacillus inoculation. J. Environ. Chem. Eng. 11, 109940 (2023).Article 

    Google Scholar 
    Wang, C. et al. Bacterial dynamics and functions driven by a novel microbial agent to promote kitchen waste composting and reduce environmental burden. J. Clean. Prod. 337, 130491 (2022).Article 

    Google Scholar 
    Wang, J., Liu, Z., Xia, J. & Chen, Y. Effect of microbial inoculation on physicochemical properties and bacterial community structure of citrus Peel composting. Bioresour Technol. 291, 121843 (2019).Article 
    PubMed 

    Google Scholar 
    Zhou, G., Xu, X., Qiu, X. & Zhang, J. Biochar influences the succession of microbial communities and the metabolic functions during rice straw composting with pig manure. Bioresour Technol. 272, 10–18 (2019).Article 
    PubMed 

    Google Scholar 
    Tortosa, G. et al. Assessment of the diversity and abundance of the total and active fungal population and its correlation with humification during two-phase Olive mill waste (alperujo) composting. Bioresour Technol. 295, 122267 (2020).Article 
    PubMed 

    Google Scholar 
    Wang, X. et al. Insights into the functionality of fungal community during the large-scale aerobic co-composting process of swine manure and rice straw. J. Environ. Manage. 270, 110958 (2020).Article 
    PubMed 

    Google Scholar 
    Gile, G. H. & Slamovits, C. H. Transcriptomic analysis reveals evidence for a cryptic plastid in the colpodellid Voromonas pontica, a close relative of Chromerids and apicomplexan parasites. PLoS One. 9, e96258 (2014).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Li, Y., Wang, Y., Zhang, S., Maurer-Alcalá, X. X. & Yan, Y. How ciliated protists survive by cysts: some key points during encystment and excystment. Front. Microbiol. 13, 785502 (2022).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Steel, H. et al. Nematode communities and macronutrients in composts and compost-amended soils as affected by feedstock composition. Appl. Soil. Ecol. 61, 100–112 (2012).Article 

    Google Scholar 
    Piedra-Buena, A. et al. Use of pepper crop residues for the control of root-knot nematodes. Bioresour Technol. 98, 2846–2851 (2007).Article 
    PubMed 

    Google Scholar 
    Siles, J. A., García-Sánchez, M. & Gómez-Brandón, M. Studying microbial communities through co-occurrence network analyses during processes of waste treatment and in organically amended soils: a review. Microorganisms 9, 1165 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Lopez, D. N., Camus, P. A., Valdivia, N. & Estay, S. A. High Temporal variability in the occurrence of consumer–resource interactions in ecological networks. Oikos 126, 1699–1707 (2017).Article 
    ADS 

    Google Scholar 
    Berry, D. & Widder, S. Deciphering microbial interactions and detecting keystone species with co-occurrence networks. Front. Microbiol. 5, 219 (2014).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Grandy, A. S., Wieder, W. R., Wickings, K. & Kyker-Snowman, E. Beyond microbes: are fauna the next frontier in soil biogeochemical models? Soil. Biol. Biochem. 102, 40–44 (2016).Article 

    Google Scholar 
    Zhang, B., Zhang, J., Liu, Y., Shi, P. & Wei, G. Co-occurrence patterns of soybean rhizosphere Microbiome at a continental scale. Soil. Biol. Biochem. 118, 178–186 (2018).Article 

    Google Scholar 
    Guo, B. et al. Microbial co-occurrence network topological properties link with reactor parameters and reveal importance of low-abundance genera. NPJ Biofilms Microbi. 8, 3 (2022).Article 

    Google Scholar 
    Gavande, P. et al. Functional characterization of thermotolerant microbial consortium for lignocellulolytic enzymes with central role of Firmicutes in rice straw depolymerization. Sci. Rep. 11, 3032 (2021).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    López, M. J. et al. Characterization of thermophilic lignocellulolytic microorganisms in composting. Front. Microbiol. 12, 697480 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wang, X. et al. Effects of inoculation of thermotolerant Bacillus strains on lignocellulose degradation. Agriculture 14 (11), 2044 (2024).Article 

    Google Scholar 
    Shi, G. et al. Interaction between nematodes and bacteria enhances soil carbon sequestration under organic material amendments. Front Microbiol 14, (2023).Download referencesFundingThis work was supported by the National Natural Science Foundation of China (32271728 and 31870501), Hubei Provincial Natural Science Foundation (2024AFB389), Youth Science Fund of Hubei Academy of Agricultural Sciences (2024NKYJJ11) and Open Project of Key Laboratory of Ministry of Agriculture and Rural Affairs (KLTMCUAR2024-03).Author informationAuthors and AffiliationsKey Laboratory of Fertilization from Agricultural Wastes, Ministry of Agriculture and Rural Affairs, Institute of Plant Protection and Soil Fertilizer, Hubei Academy of Agricultural Sciences, Wuhan, 430064, ChinaMin Xu, Yabin Zhan, Hao Fan & Yunfeng ChenNational Key Laboratory of Agricultural Microbiology, Huazhong Agricultural University, Wuhan, 430070, ChinaMin XuDepartment of Environmental Science and Engineering, College of Resources and Environmental Sciences, Nanjing Agricultural University, Nanjing, 210095, ChinaJingang XuAuthorsMin XuView author publicationsSearch author on:PubMed Google ScholarYabin ZhanView author publicationsSearch author on:PubMed Google ScholarJingang XuView author publicationsSearch author on:PubMed Google ScholarHao FanView author publicationsSearch author on:PubMed Google ScholarYunfeng ChenView author publicationsSearch author on:PubMed Google ScholarContributionsM.X. designed the study, performed the data analyses, prepared all figures, and wrote the main manuscript. J.X. and H.F. conducted the experiments and measurements. Y.Z. contributed to parts of the manuscript writing and provided scientific guidance. Y.C. acquired funding and supervised the study. All authors reviewed the manuscript.Corresponding authorCorrespondence to
    Yunfeng Chen.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 articleXu, M., Zhan, Y., Xu, J. et al. Succession of micro-food webs and their role in lignocellulose degradation during pepper stalk composting.
    Sci Rep 15, 44922 (2025). https://doi.org/10.1038/s41598-025-29221-wDownload citationReceived: 05 July 2025Accepted: 14 November 2025Published: 29 December 2025Version of record: 29 December 2025DOI: https://doi.org/10.1038/s41598-025-29221-wShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative
    KeywordsAerobic compostPhagotrophic protistNematodeMultitrophic diversityMicro-food web complexity More

  • in

    Assessing eco-environmental quality and its drivers in the Shandong section of the Yellow River Basin with an improved remote sensing ecological index

    AbstractThe Shandong section of the Yellow River Basin (SDYRB), a critical zone for ecological security in the lower reaches of the Yellow River, faces multiple ecological challenges including salinization, soil erosion, water scarcity, and anthropogenic pollution. These issues significantly hinder regional sustainable development. To assess eco-environmental quality in the SDYRB accurately, an Improved Remote Sensing Ecological Index (IRSEI) was developed by integrating the Composite Salinity Index (CSI) and Soil–Water Conservation Function Index (SWCFI). Utilizing multi-temporal imagery (2009–2023), this study analyzed spatio-temporal patterns of eco-environmental quality and their driving mechanisms. The results show that: (1) The overall eco-environmental quality exhibits a declining trend, with a spatial distribution pattern characterized as “superior in the west and poorer in the east”. High-quality areas were concentrated in western plains and Yellow River riparian zones, versus low-quality areas in eastern/northern coasts. (2) The global Moran’s I approached 1 and exhibited a gradual year-by-year decline, indicating persistent spatial agglomeration of ecological quality. Local spatial autocorrelation was predominantly characterized by High-High (H–H) and Low-Low (L–L) agglomerations, with low-value areas exhibiting an outward spread tendency. (3) Ecological quality fluctuated, declining significantly (2009–2014) before recovering (2019–2023). Degradation hotspots were identified in the northeast and southwest, whereas the improved areas were concentrated in the central region. (4) Ordinary Least Squares (OLS) regression and GeoDetector (GD) identified synergistic natural and anthropogenic driving factors: mean annual temperature, evapotranspiration, nighttime light intensity, and land use were dominant. This study improves the applicability and interpretability of IRSEI in salinized and soil-eroded regions by integrating CSI and SWCFI, offering a scientific foundation for ecological conservation and high-quality development in the SDYRB. The approach can also be extended to dynamic monitoring and evaluation of other similarly vulnerable ecological zones.

    Data availability

    The authors confirm that the data supporting the findings of this study are available within the article.
    ReferencesShi, H. et al. Spatio-temporal pattern of urban-rural income gap in the Yellow River Basin and its response to urbanization. Arid Land Geogr. 47, 1781–1793 (2024).
    Google Scholar 
    Zhang, C.; Chen, Y. Research on the protection and development of historical and cultural resources in the Yellow River Basin from the perspective of ecological civilization. 3C TIC. Cuadernos de desar rollo aplicados a las TIC 2023, 2, 156–171.Li, D., Xu, G., Tang, J., Chao, Xu. & Xia, R. Countermeasures of the Eco-environment Governance in the Lower Reach of the Yellow River. Environ. Prot. 51, 23–26 (2024).
    Google Scholar 
    Jiang, C., Zhang, H., Wang, X., Feng, Y. & Labzovskii, L. Challenging the land degradation in China’s Loess Plateau: Benefits, limitations, sustainability, and adaptive strategies of soil and water conservation. Ecol. Eng. 127, 135–150 (2019).
    Google Scholar 
    Fu, B. Ecological and environmental effects of land-use changes in the Loess Plateau of China (in Chinese). Chin. Sci. Bull. 67, 3768–3779 (2022).
    Google Scholar 
    Han, X., Gao, Q. & Yuan, B. Analysis and identification of pollution sources of comprehensive river water quality: Evidence from two river basins in China. Ecol. Indic. 135, 108561 (2022).
    Google Scholar 
    Chen, C.; Liu, R. On regional collaborative legislation in the governance of the Yellow River Basin. Proceedings of the 2021 National Symposium on Environmental and Resource Law, Taiyuan, China, 15–17 October 2021.Outline of the Ecological Protection and High-Quality Development Plan for the Yellow River Basin. Available online: http://www.gov.cn/zhengce/2021-10/08/content_5641438.htm (accessed on 8 October 2021).Ma, H., Lu, N., Jiang, Y., Huang, Q. & Wu, F. Study on green farmland construction and agricultural high quality development model in the Yellow River Basin of Shandong Province. Agric. Compr. Dev. China 06, 16–21 (2023).
    Google Scholar 
    Zhang, C. & Wang, G. Thoughts on ecological protection and high quality development of the Yellow River Basin. Yellow River 46, 1–7 (2024).
    Google Scholar 
    Nagendra, H. et al. Remote sensing for conservation monitoring: Assessing protected areas, habitat extent, habitat condition, species diversity, and threats. Ecol. Indic. 33, 45–59 (2013).
    Google Scholar 
    Hu, X. & Xu, H. A new remote sensing index for assessing the spatial heterogeneity in urban ecological quality: A case from Fuzhou City. China. Ecol. Indic. 89, 11–21 (2018).
    Google Scholar 
    Li, S., Li, C. & Kang, X. Development status and future prospects of multi-source remote sensing image fusion. Natl. Remote Sens. Bull. 25, 148–166 (2021).
    Google Scholar 
    Zhang, Y. et al. Opportunities and challenges in remote sensing applications to ecosystem ecology. Chin. J. Ecol. 36, 809–823 (2017).
    Google Scholar 
    Rumora, L., Majić, I., Miler, M. & Medak, D. Spatial video remote sensing for urban vegetation mapping using vegetation indices. Urban Ecosystems. Urban Ecosyst. 24, 1–13 (2020).
    Google Scholar 
    Xu, H. A study on information extraction of water body with the modified normalized difference water index (MNDWI). Natl. Remote Sens. Bull. 9, 589–595 (2005).
    Google Scholar 
    Xu, H. A remote sensing index for assessment of regional ecological changes. China Environ. Sci. 33, 889–897 (2013).
    Google Scholar 
    Yan, Z. Assessment of ecological environment quality in Jiawang district based on remote sensing ecological index. Master’s Thesis, China University of Mining and Technology, Xuzhou, China, May 2023.Li, R. Study on the temporal and spatial changes of the ecological environment quality of the Shendong mining area in Shanxi, Shaanxi, and Inner Mongolia. Master’s Thesis, Inner Mongolia Agricultural University, Hohhot, China, June 2021.Zhang, Q., Wang, W. & Liu, J. Remote sensing monitoring and change analysis of ecological environment in Gulang County. Trop. Geomorphol. 42, 1–8 (2021).
    Google Scholar 
    Li, H., Huang, J., Liang, Y., Wang, H. & Zhang, Y. Evaluating the quality of ecological environment in Wuhan based on remote sensing ecological index. J. Yunnan Univ. Nat. Sci. Ed. 42, 81–90 (2020).
    Google Scholar 
    Kaixuan, Y. et al. Spatiotemporal changes of eco-environmental quality based on remote sensing-based ecological index in the Hotan Oasis, Xinjiang. J. Arid Land 03, 262–283 (2022).
    Google Scholar 
    Yang, X., Shi, Y., Feng, S., Wang, C. & Chai, J. Major ecological environment problems and countermeasures in the downstream of the Yellow River (Shandong Section). Shandong Land Resour. 26(01), 15–18 (2010).
    Google Scholar 
    Pang, Q. et al. Problems and Countermeasures of soil and water loss control in the Loess Plateau in the new period. Yellow River 44(S1), 73–74 (2022).
    Google Scholar 
    Dong, X. et al. Spatiotemporal evolution and influencing mechanism of urbanization and ecological environmental quality between 2000 and 2020 in Henan Province, China. Remote Sens. Appl. Soc. Environ. 37, 101492 (2025).
    Google Scholar 
    Xin, J., Yang, J., Yu, H., Ren, J. & Yu, W. Towards ecological civilization: Spatiotemporal heterogeneity and drivers of ecological quality transitions in China (2001–2020). Appl. Geogr. 173, 103439 (2024).
    Google Scholar 
    Liu, Q., Cai, J., Xie, M., Tian, J. & He, W. Assessment of ecosystem service value of national wetland parks along the Shandong section of the Yellow River. Wetl. Sci. 22, 697–706 (2024).
    Google Scholar 
    Zhou, C., Zhao, Y., Ren, M., Jin, Y. & Lu, S. Spatiotemporal differentiation, center of gravity evolution and driving factors of national wetland parks in the Yellow River Basin. Arid Land Geogr. 47, 506–514 (2024).
    Google Scholar 
    Zhang, Y. Effects of oil pollution and Spartina alterniflora invasion on Phragmites australis communities in the Yellow River Delta. Master’s Thesis, Shandong University, Qingdao, China, June 2023.Luo, P., Zhu, W. & Wang, S. Consideration on comprehensive legislation of high-quality development in the Yellow River Basin. People Yellow River 45, 14–18 (2023).
    Google Scholar 
    Lu, D. & Sun, D. Comprehensive Management and Sustainable Development of the Yellow River Basin. Acta Geogr. Sin. 74, 2431–2436 (2019).
    Google Scholar 
    Li, Y., Wang, Y. & Song, K. Analysis of spatiotemporal changes and driving forces of eco⁃ logical quality in the main stream of the Yangtze River (Jiangsu Section) from 2000 to 2020 based on RSEI and Geo Detector. Remote Sens. Technol. Appl. 39, 1478–1489 (2024).
    Google Scholar 
    Fu, K., Jia, G., Yu, X. & Wang, X. Ecological environment assessment and driving mechanism analysis of Nagqu and Amdo sections of Qinghai-Xizang highway based on improved remote sensing ecological index. Environ. Sci. 45, 1586–1597 (2024).
    Google Scholar 
    Gao, Y., Liu, Y. & Xu, H. Analysis of the spatio-temporal evolution and driving factors of urban ecological quality based on long-term Landsat image time series. Acta Geod. Cartogr. Sin. 54, 510–522 (2025).
    Google Scholar 
    Quan, W., Zhou, H. & Wang, W. Characteristics of ecological environment quality dynamic of the urban agglomeration in Guanzhong plain from 2000 to 2023. Res. Soil Water Conserv. 32, 336–346 (2025).
    Google Scholar 
    Duo, L. et al. Assessing the spatiotemporal evolution and drivers of ecological environment quality using an enhanced remote sensing ecological index in Lanzhou City. China. Remote Sens. 15, 4704 (2023).
    Google Scholar 
    Ji, J. et al. Spatiotemporal and multiscale analysis of the coupling coordination degree between economic development equality and eco-environmental quality in China from 2001 to 2020. Remote Sens. 14, 737 (2022).
    Google Scholar 
    Xiong, Y. et al. Assessment of spatial–temporal changes of ecological environment quality based on RSEI and GEE: A case study in Erhai Lake Basin, Yunnan province. China. Ecol. Indic. 125, 107518 (2021).
    Google Scholar 
    Xu, H., Li, C. & Lin, M. Should RSEI Use PCA or kPCA?. Geomat. Inf. Sci. Wuhan Univ. 4, 506–513 (2023).
    Google Scholar 
    Liu, Z., Xu, H., Li, L., Tang, F. & Lin, Z. Ecological change in the Hangzhou area using the remote sensing based ecological index. J. Basic Sci. Eng. 23, 728–739 (2015).
    Google Scholar 
    Luo, Q., Liu, F. & Zhu, K. Spatiotemporal differentiation and spatial autocorrelation analysis of landscape ecological risk in the Chishui River Basin (Guizhou section). Res. Soil Water Conserv. 32, 282–290 (2025).
    Google Scholar 
    Shao, Z., Zhang, F., Peng, K. & Zhang, F. Ecological environment quality monitoring of Qitai oasis based on land use/cover change and remote sensing ecological index. Environ. Sci. 45, 5890–5899 (2024).
    Google Scholar 
    Tian, K. et al. Spatio-temporal evolution characteristics of cultivated land and its driving force in Karst mountainous area: A case study of Bijie City. Guizhou Agric. Sci. 51, 113–125 (2023).
    Google Scholar 
    Xiong, Y. et al. Assessment of spatial-temporal changes of ecological environment quality based on RSEI and GEE: A case study in Erhai Lake Basin, Yunnan Province, China. Ecol. Indic. 125, 107518 (2021).
    Google Scholar 
    Chen, Z., Wang, X. & Guo, J. Analysis of spatial and temporal variations and driving factors of ecological environment quality in Huangshi city. Bull. Surv. Mapp. 04, 145–151 (2025).
    Google Scholar 
    Liu, M. Geometric analysis of ordinary least square method. Stat. Decis. 04, 90–92 (2012).
    Google Scholar 
    Wang, J. & Xu, C. Geo detector: Principle and prospect. Acta Geogr. Sin. 72, 116–134 (2017).
    Google Scholar 
    Su, Y. et al. Pervasive but biome-dependent relationship between fragmentation and resilience in forests. Nat. Ecol. Evol. 9, 1670–1684 (2025).
    Google Scholar 
    Liu, X. et al. Spatial patterns of soil polycyclic aromatic hydrocarbon contamination and source-specific stratified probabilistic health risk assessment at a decommissioned coking plant in North China. J. Environ. Chem. Eng. 13, 120242 (2025).
    Google Scholar 
    Wu, Q. et al. Tracing threats across the land–marine transition areas: Decoding heavy metal sources and risks in soil-water-sediment systems of northern and eastern coastal China. Mar. Pollut. Bull. 223, 118969 (2026).
    Google Scholar 
    Wang, S. et al. Anthropogenic impacts on the Yellow River Basin. Nat. Rev. Earth Environ. 6, 656–671 (2025).
    Google Scholar 
    Ma, X. Construction and improvement of water pollution prevention mechanism: Comment on “water pollution prevention action plan”. Yellow River 43(05), 165 (2021).
    Google Scholar 
    Li, Y. On systematic governance of mountain-river-forest-Farmland-Lake-Grassland-Sand in the Yellow River Basin: Logic and paths. Adm. Reform 03, 56–64 (2025).
    Google Scholar 
    Chen, D. & Huang, Q. The new policy for innovative transformation in regional industrial chains, the conversion of new and old kinetic energy, and energy poverty alleviation. Energies 17, 2667 (2024).
    Google Scholar 
    Zheng, Y., Ge, C. & Yu, F. Implementation mechanism of agricultural intensification, greening and resource-recycling: From a Low-carbon perspective. J. Huazhong Agric. Univ. (Soc. Sci. Ed.) 01, 32–44 (2022).
    Google Scholar 
    Zhou, G., Zhai, X. & Zhang, S. Study on the spatial-temporal pattern and influencing factors of agricultural green development in the Yellow River Basin. Yellow River 47(04), 7–14 (2025).
    Google Scholar 
    Zhao, K. & Ma, R. Research on the integration of urban and rural eco-environments under the contexts of industrialization and urbanization a case study of Taiyuan City of Shanxi Province. J. North Univ. China (Soc. Sci. Ed.) 34(02), 51–58 (2018).
    Google Scholar 
    Zhu, T., Yao, W. & Zhu, L. Research on the coupling between industrialization and ecotope. Environ. Pollut. Control 36(06), 78–83 (2014).
    Google Scholar 
    Yao, G. et al. The coupling coordination degree of ecological environment and urbanization in ecological civilization demonstration area based on remote sensing ecological index and compounded night light index: A case study of the Yangtze River delta. J. Ecol. Rural Environ. 39, 1386–1398 (2023).
    Google Scholar 
    Zhang, J. & Yang, Y. Remote sensing evaluation on the change of ecological status of Pearl River delta urban agglomeration. J. Northwest For. Univ. 34, 184–191 (2019).
    Google Scholar 
    Martín-López, B. et al. Delineating boundaries of social-ecological systems for landscape planning: A comprehensive spatial approach. Land Use Policy 66, 90–104 (2017).
    Google Scholar 
    Xu, H. & Yang, S. Urban heat island effect and urban ecosystems. J. Beijing Norm. Univ. (Nat. Sci.) 54(06), 790–798 (2018).
    Google Scholar 
    Bai, Y. et al. The coupled effect of soil and atmospheric constraints on the vulnerability and water use of two desert riparian ecosystems. Agric. For. Meteorol. 311, 108701 (2021).
    Google Scholar 
    Dehghan Rahimabadi, P. et al. The Nexus between land use/cover changes and land surface temperature: Remote sensing based two-decadal analysis. J. Arid Environ. 225, 105269 (2024).
    Google Scholar 
    Zhang, H. et al. An empirical analysis of tourism eco-efficiency in ecological protection priority areas based on the DPSIR-SBM model: A case study of the Yellow River Basin, China. Ecol. Inform. 70, 101720 (2022).
    Google Scholar 
    Hu, B. et al. Coupling strength of human-natural systems mediates the response of ecosystem services to land use change. J. Environ. Manage. 344, 118521 (2023).
    Google Scholar 
    Wang, Y. et al. Evaluating the ecological conservation effectiveness and strategies in Nanling key ecological function zone of China: A CHANS perspective. Ecol. Front. 44(6), 1140–1148 (2024).
    Google Scholar 
    Download referencesFundingThe research was funded by the Jinan Municipal School-Integration Development Strategy Program (Phase II) (Grant No. JNSX2023107), Science and Technology Program of Shandong Provincial Department of Housing and Urban–Rural Development (Grant No. 2025KYKF-CSAQ182), Research Program of Qilu Institute of Technology (Grant No. QIT24TP006 and QIT24NN085) and Shandong Provincial Natural Science Foundation (Grant No. ZR2024QE385).Author informationAuthors and AffiliationsCivil Engineering Department, Qilu Institute of Technology, Jinan, 250200, ChinaPeipei Wang, Chun-Pin Tseng, Qinghao Wei, Min Qiao, Xiaoshuang Li & Ran AnMathematics Department, Brandeis University, Waltham, MA, USAYiyou FanAuthorsPeipei WangView author publicationsSearch author on:PubMed Google ScholarChun-Pin TsengView author publicationsSearch author on:PubMed Google ScholarYiyou FanView author publicationsSearch author on:PubMed Google ScholarQinghao WeiView author publicationsSearch author on:PubMed Google ScholarMin QiaoView author publicationsSearch author on:PubMed Google ScholarXiaoshuang LiView author publicationsSearch author on:PubMed Google ScholarRan AnView author publicationsSearch author on:PubMed Google ScholarContributionsConceptualization, P.W. and Q.W.; methodology, P.W. and C.T.; software, P.W.; validation, P.W., C.T. and P.W.; formal analysis, Q.W. and M.Q.; investigation, Y.F.; resources, P.W.; data curation, P.W. and Q.W.; writing—original draft preparation, P.W. and Q.W.; writing—review and editing, P.W. and C.T.; visualization, C.T., and M.Q.; supervision, X.L., and R.A.; project administration, R.A.; funding acquisition, C.T. All authors have read and agreed to the published version of the manuscript.Corresponding authorCorrespondence to
    Chun-Pin Tseng.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.Rights 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 articleWang, P., Tseng, CP., Fan, Y. et al. Assessing eco-environmental quality and its drivers in the Shandong section of the Yellow River Basin with an improved remote sensing ecological index.
    Sci Rep (2025). https://doi.org/10.1038/s41598-025-31580-3Download citationReceived: 16 July 2025Accepted: 03 December 2025Published: 28 December 2025DOI: https://doi.org/10.1038/s41598-025-31580-3Share 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
    KeywordsThe Shandong section of the Yellow River Basin (SDYRB)Eco-environmental qualityImproved remote sensing ecological index (IRSEI)Spatial autocorrelationDriving factors More