More stories

  • in

    Detection of energetic equivalence depends on food web architecture and estimators of energy use

    AbstractEcologists have long debated the universality of the energetic equivalence rule, which posits that population energy use should be invariant with average body size due to negative size–density scaling. We explore size–density and size–energy use scaling across 183 geographically–distributed soil invertebrate food webs (comprising 55,054 individual soil invertebrates) to investigate the universality of these fundamental energetic equivalence rule assumptions across trophic levels and varying food web structure. Additionally, we compare two measures of energy use to investigate size–energy use relationships: population metabolism and energy fluxes. We find that size–density scaling does not support energetic equivalence in soil communities. Furthermore, evidence of energetic equivalence is dependent on the estimate of energy use applied, the trophic level of consumers, and food web properties. Our study demonstrates a need to integrate food web energetics and trophic structure to better understand how energetic constraints shape the body size structure of terrestrial ecosystems.

    Similar content being viewed by others

    A species-level multi-trophic metaweb for Switzerland

    Article
    Open access
    09 July 2025

    The productivity–stability trade-off in global food systems

    Article

    03 September 2024

    A network simplification approach to ease topological studies about the food-web architecture

    Article
    Open access
    17 August 2022

    Data availability

    The data generated in this study has been deposited in the figshare repository https://doi.org/10.6084/m9.figshare.25591254.v1. The raw EFForTS and ECOWORM data are protected and are not available due to data privacy laws. Source data are provided with this paper.
    Code availability

    The code generated in this study has been deposited in the figshare repository https://doi.org/10.6084/m9.figshare.25591227.v1.
    ReferencesHatton, I. A., Dobson, A. P., Storch, D., Galbraith, E. D. & Loreau, M. Linking scaling laws across eukaryotes. Proc. Natl. Acad. Sci. Usa. 116, 21616–21622 (2019).
    Google Scholar 
    Cohen, J. E., Jonsson, T. & Carpenter, S. R. Ecological community description using the food web, species abundance, and body size. Proc. Natl. Acad. Sci. USA 100, 1781–1786 (2003).
    Google Scholar 
    Woodward, G. et al. Body size in ecological networks. Trends Ecol. Evol. 20, 402–409 (2005).
    Google Scholar 
    Brown, J. H., Gillooly, J. F., Allen, A. P., Savage, V. M. & West, G. B. Toward a metabolic theory of ecology. Ecology 85, 1771–1789 (2004).
    Google Scholar 
    Damuth, J. Interspecific allometry of population density in mammals and other animals: the independence of body mass and population energy-use. Biol. J. Linn. Soc. 31, 193–246 (1987).
    Google Scholar 
    Meehan, T. D. Energy use and animal abundance in litter and soil communities. Ecology 87, 1650–1658 (2006).
    Google Scholar 
    Meehan, T. D. Mass and temperature dependence of metabolic rate in litter and soil invertebrates. Physiol. Biochem. Zool. 79, 878–884 (2006).
    Google Scholar 
    Ott, D. et al. Litter elemental stoichiometry and biomass densities of forest soil invertebrates. Oikos 123, 1212–1223 (2014).
    Google Scholar 
    Ott, D. et al. Unifying elemental stoichiometry and metabolic theory in predicting species abundances. Ecol. Lett. 17, 1247–1256 (2014).
    Google Scholar 
    Antunes, A. C. et al. Environmental drivers of local abundance–mass scaling in soil animal communities. Oikos 2023, e09735 (2023).
    Google Scholar 
    Mulder, C. & Elser, J. J. Soil acidity, ecological stoichiometry and allometric scaling in grassland food webs. Glob. Change Biol. 15, 2730–2738 (2009).
    Google Scholar 
    Cyr, H., Downing, J. A., Peters, R. H. & Cyr, H. Density-body size relationships in local aquatic communities. Oikos 79, 333 (1997).
    Google Scholar 
    Cyr, H., Peters, R. H., Downing, J. A. & Cyr, H. Population density and community xize structure: comparison of aquatic and terrestrial systems. Oikos 80, 139 (1997).
    Google Scholar 
    White, E. P., Ernest, S. K. M., Kerkhoff, A. J. & Enquist, B. J. Relationships between body size and abundance in ecology. Trends Ecol. Evol. 22, 323–330 (2007).
    Google Scholar 
    Ehnes, R. B. et al. Lack of energetic equivalence in forest soil invertebrates. Ecology 95, 527–537 (2014).
    Google Scholar 
    Damuth, J. Population density and body size in mammals. Nature 290, 699–700 (1981).
    Google Scholar 
    Meehan, T. D. et al. Energetic equivalence in a soil arthropod community from an aspen–conifer forest. Pedobiologia 50, 307–312 (2006).
    Google Scholar 
    Trebilco, R., Baum, J. K., Salomon, A. K. & Dulvy, N. K. Ecosystem ecology: size-based constraints on the pyramids of life. Trends Ecol. Evol. 28, 423–431 (2013).
    Google Scholar 
    Reuman, D. C., Mulder, C., Raffaelli, D. & Cohen, J. E. Three allometric relations of population density to body mass: theoretical integration and empirical tests in 149 food webs. Ecol. Lett. 11, 1216–1228 (2008).
    Google Scholar 
    Reuman, D. C. et al. Chapter 1 allometry of body size and abundance in 166 food webs. in Advances in Ecological Research vol. 41 1–44 (Elsevier, 2009).Arim, M., Abades, S. R., Laufer, G., Loureiro, M. & Marquet, P. A. Food web structure and body size: trophic position and resource acquisition. Oikos 119, 147–153 (2010).
    Google Scholar 
    Gjoni, V. & Glazier, D. S. A perspective on body size and abundance relationships across ecological communities. Biology 9, 42 (2020).
    Google Scholar 
    Mulder, C., Vonk, J. A., Den Hollander, H. A., Hendriks, A. J. & Breure, A. M. How allometric scaling relates to soil abiotics. Oikos 120, 529–536 (2011).
    Google Scholar 
    Gjoni, V., Marle, P., Ibelings, B. W. & Castella, E. Size–abundance relationships of freshwater macroinvertebrates in two contrasting floodplain channels of rhone river. Water 14, 794 (2022).
    Google Scholar 
    Gjoni, V., Glazier, D. S., Wesner, J. S., Ibelings, B. W. & Thomas, M. K. Temperature, resources and predation interact to shape phytoplankton size–abundance relationships at a continental scale. Glob. Ecol. Biogeogr. 32, 2006–2016 (2023).
    Google Scholar 
    Gjoni, V., Cozzoli, F., Rosati, I. & Basset, A. Size–density relationships: a cross-community approach to benthic macroinvertebrates in mediterranean and black sea lagoons. Estuaries Coasts 40, 1142–1158 (2017).
    Google Scholar 
    Loeuille, N. & Loreau, M. Evolution of body size in food webs: does the energetic equivalence rule hold? Ecol. Lett. 9, 171–178 (2006).
    Google Scholar 
    Barnes, A. D. et al. Energy flux: the link between multitrophic biodiversity and ecosystem functioning. Trends Ecol. Evol. 33, 186–197 (2018).
    Google Scholar 
    Gauzens, B. et al. Fluxweb: package to easily estimate energy fluxes in food webs. Methods Ecol. Evol. 10, 270–279 (2019).
    Google Scholar 
    Rooney, N., McCann, K. S. & Moore, J. C. A landscape theory for food web architecture. Ecol. Lett. 11, 867–881 (2008).
    Google Scholar 
    Polis, G. A. & Strong, D. R. Food web complexity and community dynamics. Am. Naturalist 147, 813–846 (1996).
    Google Scholar 
    Eisenhauer, N. Aboveground–belowground interactions as a source of complementarity effects in biodiversity experiments. Plant Soil 351, 1–22 (2012).
    Google Scholar 
    Lembrechts, J. J. et al. Global maps of soil temperature. Glob. Change Biol. 28, 3110–3144 (2022).
    Google Scholar 
    Potapov, A. M. et al. Size compartmentalization of energy channeling in terrestrial belowground food webs. Ecology 102, e03421 (2021).
    Google Scholar 
    Glazier, D. S. Variable metabolic scaling breaks the law: from ‘Newtonian’ to ‘Darwinian’ approaches. Proc. R. Soc. B. 289, 20221605 (2022).
    Google Scholar 
    Romera, P. J. et al. Ecological succession shapes size–density scaling relationships of trees and soil invertebrates. Funct. Ecol. 38, 2156–2168 (2024).
    Google Scholar 
    Potapov, A. M., Klarner, B., Sandmann, D., Widyastuti, R. & Scheu, S. Linking size spectrum, energy flux and trophic multifunctionality in soil food webs of tropical land-use systems. J. Anim. Ecol. 88, 1845–1859 (2019).
    Google Scholar 
    Damuth, J. A macroevolutionary explanation for energy equivalence in the scaling of body size and population density. Am. Naturalist 169, 621–631 (2007).
    Google Scholar 
    Poisot, T., Mouquet, N. & Gravel, D. Trophic complementarity drives the biodiversity–ecosystem functioning relationship in food webs. Ecol. Lett. 16, 853–861 (2013).
    Google Scholar 
    Ulrich, W. et al. Temporal patterns of energy equivalence in temperate soil invertebrates. Oecologia 179, 271–280 (2015).
    Google Scholar 
    Drescher, J. et al. Ecological and socio-economic functions across tropical land use systems after rainforest conversion. Philos. Trans. R. Soc. B 371, 20150275 (2016).
    Google Scholar 
    Jochum, M. et al. Earthworm invasion causes declines across soil fauna size classes and biodiversity facets in northern North American forests. Oikos 130, 766–780 (2021).
    Google Scholar 
    Naumann, I. D. The Insects of Australia: A Textbook for Students and Research Workers: Free Download, Borrow, and Streaming: Internet Archive. (Melbourne University Press, 1991).Barnes, A. D. et al. Consequences of tropical land use for multitrophic biodiversity and ecosystem functioning. Nat. Commun. 5, 5351 (2014).
    Google Scholar 
    Mercer, R. D., Gabriel, A. G. A., Barendse, J., Marshall, D. J. & Chown, S. L. Invertebrate body sizes from Marion Island. Antartic Sci. 13, 135–143 (2001).
    Google Scholar 
    Sohlström, E. H. et al. Applying generalized allometric regressions to predict live body mass of tropical and temperate arthropods. Ecol. Evol. 8, 12737–12749 (2018).
    Google Scholar 
    Ehnes, R. B., Rall, B. C. & Brose, U. Phylogenetic grouping, curvature and metabolic scaling in terrestrial invertebrates: invertebrate metabolism. Ecol. Lett. 14, 993–1000 (2011).
    Google Scholar 
    Potapov, A. M. et al. Feeding habits and multifunctional classification of soil-associated consumers from protists to vertebrates. Biol. Rev. 97, 1057–1117 (2022).
    Google Scholar 
    Barnes, A. D. et al. Biodiversity enhances the multitrophic control of arthropod herbivory. Sci. Adv. 6, eabb6603 (2020).
    Google Scholar 
    Lang, B., Ehnes, R. B., Brose, U. & Rall, B. C. Temperature and consumer type dependencies of energy flows in natural communities. Oikos 126, 1717–1725 (2017).
    Google Scholar 
    De Ruiter, P. C., Van Veen, J. A., Moore, J. C., Brussaard, L. & Hunt, H. W. Calculation of nitrogen mineralization in soil food webs. Plant Soil 157, 263–273 (1993).
    Google Scholar 
    Perkins, D. M. et al. Energetic equivalence underpins the size structure of tree and phytoplankton communities. Nat. Commun. 10, 255 (2019).
    Google Scholar 
    Edwards, A. M., Robinson, J. P. W., Plank, M. J., Baum, J. K. & Blanchard, J. L. Testing and recommending methods for fitting size spectra to data. Methods Ecol. Evol. 8, 57–67 (2017).
    Google Scholar 
    R Core Team. R: a language and environment for statistical computing. R Foundation for Statistical Computing (R Core Team, 2023).Legendre, P. Model II regression user’s guide, R edition. chrome-extension://efaidnbmnnnibpcajpcglclefindmkaj/https://cran.r-project.org/web/packages/lmodel2/vignettes/mod2user.pdf (2018)Download referencesAcknowledgementsOur project was supported by the Marsden Fund Council from Government funding managed by Royal Society Te Apārangi (grant MFP-23-UOW-029), and the People, Cities, and Nature research programme (Ministry of Business, Innovation and Employment, grant UOWX2101). We thank the numerous people that assisted in the field and laboratory and mana whenua (Indigenous people) of the land our sites were on. We acknowledge the use of data drawn from the EFForTS and ECOWORM projects. All authors gratefully acknowledge the support of iDiv, which is funded by the German Research Foundation (DFG – FZT 118, 202548816). N.E. and O.F. thank the DFG (Ei 862/29–1; Ei 862/31–1) for funding. Fig. 1, and Fig. 2 were created with Canva.com using images and elements of these images created by authors (see supplementary code). A.P. was funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – Projektnummer 493345801.Author informationAuthors and AffiliationsTe Aka Mātuatua – School of Science, Te Whare Wānanga o Waikato – University of Waikato, Private Bag, Hamilton, New ZealandPoppy Joaquina Romera, Bibishan Rai, Kiri Joy Wallace & Andrew D. BarnesGerman Centre of Integrative Biodiversity Research (iDiv), Halle–Jena–Leipzig, Puschstraße 4, Leipzig, GermanyBenoit Gauzens, Ana Carolina Antunes, Ulrich Brose, Nico Eisenhauer, Olga Ferlian, Myriam R. Hirt, Malte Jochum & Anton PotapovInstitute of Biodiversity, Friedrich Schiller University Jena, Jena, GermanyBenoit Gauzens, Ulrich Brose, Olga Ferlian & Myriam R. HirtInstitute of Biology, Leipzig University, Puschstraße 4, Leipzig, GermanyNico Eisenhauer & Malte JochumDepartment of Global Change Ecology, Biocentre, University of Würzburg, Emil–Hilb–Weg 22, Würzburg, GermanyMalte JochumManaaki Whenua – Landcare Research, Private Bag 3127, Waikato Mail Centre, Hamilton, New ZealandGrace MitchellCentre for Biodiversity Monitoring and Conservation Science, Leibniz–Institute for the Analysis of Biodiversity Change (LIB), Adenauerallee 127, Bonn, GermanyDavid OttSenckenberg Museum for Natural History Görlitz, Görlitz, GermanyAnton PotapovInternational Institute Zittau, TUD Dresden University of Technology, Zittau, GermanyAnton PotapovJFB Institute of Zoology and Anthropology, University of Göttingen, Untere Karspüle 2, Göttingen, GermanyStefan ScheuCentre of Biodiversity and Sustainable Land Use, University of Göttingen, Büsgenweg 1, Göttingen, GermanyStefan ScheuAuthorsPoppy Joaquina RomeraView author publicationsSearch author on:PubMed Google ScholarBenoit GauzensView author publicationsSearch author on:PubMed Google ScholarAna Carolina AntunesView author publicationsSearch author on:PubMed Google ScholarUlrich BroseView author publicationsSearch author on:PubMed Google ScholarNico EisenhauerView author publicationsSearch author on:PubMed Google ScholarOlga FerlianView author publicationsSearch author on:PubMed Google ScholarMyriam R. HirtView author publicationsSearch author on:PubMed Google ScholarMalte JochumView author publicationsSearch author on:PubMed Google ScholarGrace MitchellView author publicationsSearch author on:PubMed Google ScholarDavid OttView author publicationsSearch author on:PubMed Google ScholarAnton PotapovView author publicationsSearch author on:PubMed Google ScholarBibishan RaiView author publicationsSearch author on:PubMed Google ScholarStefan ScheuView author publicationsSearch author on:PubMed Google ScholarKiri Joy WallaceView author publicationsSearch author on:PubMed Google ScholarAndrew D. BarnesView author publicationsSearch author on:PubMed Google ScholarContributionsP.J.R. and A.D.B. conceived the study. P.J.R., A.D.B., A.P., B.R., G.M., K.J.W., M.J., S.S., D.O., M.R.H., U.B., N.E., and O.F. collected and processed the soil data. A.C.A. curated and processed the EFForTS and ECOWORM soil data. P.J.R. and B.G. analysed the data. P.J.R. and A.D.B. wrote the first draft of the manuscript, and all authors contributed substantially to revisions.Corresponding authorCorrespondence to
    Poppy Joaquina Romera.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Peer review

    Peer review information
    Nature Communications thanks Matias Arim, Douglas Glazier, and Peter de Ruiter for their contribution to the peer review of this work. A peer review file is available.

    Additional informationPublisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary informationSupplementary InformationReporting SummaryTransparent Peer Review fileSource dataSource DataRights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleRomera, P.J., Gauzens, B., Antunes, A.C. et al. Detection of energetic equivalence depends on food web architecture and estimators of energy use.
    Nat Commun (2025). https://doi.org/10.1038/s41467-025-67615-6Download citationReceived: 21 February 2025Accepted: 04 December 2025Published: 19 December 2025DOI: https://doi.org/10.1038/s41467-025-67615-6Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Global coral genomic vulnerability explains recent reef losses

    AbstractThe dramatic decline of reef-building corals calls for a better understanding of coral adaptation to ocean warming. Here, we characterize genetic diversity of the widespread genus Acropora by building a genomic database of 595 coral samples from different oceanic regions—from the Great Barrier Reef to the Persian Gulf. Through genome-environment associations, we find that different Acropora species show parallel evolutionary signals of heat-adaptation in the same genomic regions, pointing to genes associated with molecular heat shock responses and symbiosis. We then project the present and the predicted future distribution of heat-adapted genotypes across reefs worldwide. Reefs projected with low frequency of heat-adapted genotypes display higher rates of Acropora decline, indicating a potential genomic vulnerability to heat exposure. Our projections also suggest a transition where heat-adapted genotypes will spread at least until 2040. However, this transition will likely involve mass mortality of entire non-adapted populations and a consequent erosion of Acropora genetic diversity. This genetic diversity loss could hinder the capacity of Acropora to adapt to the more extreme heatwaves projected beyond 2040. Genomic vulnerability and genetic diversity loss estimates can be used to reassess which coral reefs are at risk and their conservation.

    Data availability

    Genomic data used in this study are publicly available in NCBI, for the full list of accession numbers and data links please see Supplementary Table 1. Processed data are available at Zenodo81 (https://doi.org/10.5281/zenodo.10838947). Supplementary Data 1 displays the list of the 85 genomic windows where genotype-environment associations were repeatedly found in different datasets.
    Code availability

    Code to reproduce the analysis is available at Zenodo81 (https://doi.org/10.5281/zenodo.10838947).
    ReferencesSouter, D. et al. Status of coral reefs of the world: 2020. (Australian government, United Nations Environment Program (UNEP), 2021).Spalding, M., Ravilious, C. & Green, E. World atlas of coral reefs. (University of California Press, Berkeley, USA, 2001).Dixon, G. B. et al. Genomic determinants of coral heat tolerance across latitudes. Science 348, 1460–1462 (2015).
    Google Scholar 
    Howells, E. J., Abrego, D., Meyer, E., Kirk, N. L. & Burt, J. A. Host adaptation and unexpected symbiont partners enable reef-building corals to tolerate extreme temperatures. Glob. Chang. Biol. 22, 2702–2714 (2016).
    Google Scholar 
    Selmoni, O. et al. Seascape genomics reveals candidate molecular targets of heat stress adaptation in three coral species. Mol. Ecol. 30, 1892–1906 (2021).
    Google Scholar 
    Thomas, L. et al. Spatially varying selection between habitats drives physiological shifts and local adaptation in a broadcast spawning coral on a remote atoll in Western Australia. Sci Adv 8, eabl9185 (2022).
    Google Scholar 
    Selmoni, O., Rochat, E., Lecellier, G., Berteaux-Lecellier, V. & Joost, S. Seascape genomics as a new tool to empower coral reef conservation strategies: An example on north-western Pacific Acropora digitifera. Evol. Appl. 13, 1923–1938 (2020).
    Google Scholar 
    Bay, R. A. & Palumbi, S. R. Multilocus adaptation associated with heat resistance in reef-building corals. Curr. Biol. 24, 2952–2956 (2014).
    Google Scholar 
    Fuller, Z. L. et al. Population genetics of the coral acropora millepora: Toward genomic prediction of bleaching. Science 369, eaba4674 (2020).
    Google Scholar 
    Cooke, I. et al. Genomic signatures in the coral holobiont reveal host adaptations driven by Holocene climate change and reef specific symbionts. Sci Adv 6, eabc6318 (2020).
    Google Scholar 
    Jin, Y. K. et al. Genetic markers for antioxidant capacity in a reef-building coral. Sci Adv 2, e1500842 (2016).
    Google Scholar 
    Selmoni, O., Bay, L. K., Exposito-Alonso, M. & Cleves, P. A. Finding genes and pathways that underlie coral adaptation. Trends Genet. 40, 213–227 (2024).
    Google Scholar 
    Skirving, W. et al. Coraltemp and the coral reef watch coral bleaching heat stress product suite version 3.1. Remote Sensing 12, 3856 (2020).
    Google Scholar 
    Riginos, C., Crandall, E. D., Liggins, L., Bongaerts, P. & Treml, E. A. Navigating the currents of seascape genomics: how spatial analyses can augment population genomic studies. Curr. Zool. 62, 581–601 (2016).
    Google Scholar 
    Rellstab, C., Dauphin, B. & Exposito-Alonso, M. Prospects and limitations of genomic offset in conservation management. Evol. Appl. 14, 1202–1212 (2021).
    Google Scholar 
    Rellstab, C., Gugerli, F., Eckert, A. J., Hancock, A. M. & Holderegger, R. A practical guide to environmental association analysis in landscape genomics. Mol. Ecol. 24, 4348–4370 (2015).
    Google Scholar 
    Booker, T. R., Yeaman, S. & Whitlock, M. C. Using genome scans to identify genes used repeatedly for adaptation. Evolution 77, 801–811 (2023).
    Google Scholar 
    Whiting, J. R. et al. The genetic architecture of repeated local adaptation to climate in distantly related plants. Nat. Ecol. Evol. 8, 1933–1947 (2024).
    Google Scholar 
    Torquato, F. et al. Population genetic structure of a major reef-building coral species Acropora downingi in northeastern Arabian Peninsula. Coral Reefs 41, 743–752 (2022).
    Google Scholar 
    Shinzato, C., Mungpakdee, S., Arakaki, N. & Satoh, N. Genome-wide SNP analysis explains coral diversity and recovery in the Ryukyu Archipelago. Sci. Rep. 5, 18211 (2015).
    Google Scholar 
    Drury, C. & Lirman, D. Genotype by environment interactions in coral bleaching. Proc. Biol. Sci. 288, 20210177 (2021).
    Google Scholar 
    Matz, M. V., Treml, E. A., Aglyamova, G. V. & Bay, L. K. Potential and limits for rapid genetic adaptation to warming in a Great Barrier Reef coral. PLoS Genet. 14, e1007220 (2018).
    Google Scholar 
    Sedlazeck, F. J., Rescheneder, P. & von Haeseler, A. NextGenMap: fast and accurate read mapping in highly polymorphic genomes. Bioinformatics 29, 2790–2791 (2013).
    Google Scholar 
    Korneliussen, T. S., Albrechtsen, A. & Nielsen, R. ANGSD: Analysis of next generation sequencing data. BMC Bioinformatics 15, 356 (2014). 2014 15:11–13.
    Google Scholar 
    Exposito-Alonso, M. et al. Genetic diversity loss in the anthropocene. Science 377, 1431–1435 (2022).
    Google Scholar 
    García-Urueña, R., Kitchen, S. A. & Schizas, N. V. Fine scale population structure of Acropora palmata and acropora cervicornis in the colombian caribbean. PeerJ 10, e13854 (2022).
    Google Scholar 
    van der Ven, R. M., Ratsimbazafy, H. A. & Kochzius, M. Large-scale biogeographic patterns are reflected in the genetic structure of a broadcast spawning stony coral. Coral Reefs 41, 611–624 (2022).
    Google Scholar 
    Caye, K., Jumentier, B., Lepeule, J. & François, O. LFMM 2: Fast and accurate inference of gene-environment associations in genome-wide studies. Mol. Biol. Evol. 36, 852–860 (2019).
    Google Scholar 
    Selmoni, O., Lecellier, G., Berteaux-Lecellier, V. & Joost, S. The reef environment centralized information system (RECIFS): An integrated geo-environmental database for coral reef research and conservation. Glob. Ecol. Biogeogr. 32, 622–632 (2023).
    Google Scholar 
    Liu, G., Strong, A. E. & Skirving, W. Remote sensing of sea surface temperatures during 2002 Barrier Reef coral bleaching. Eos Trans. Amer. Geophys. Union 84, 137–141 (2003).
    Google Scholar 
    Selmoni, O., Vajana, E., Guillaume, A., Rochat, E. & Joost, S. Sampling strategy optimization to increase statistical power in landscape genomics: A simulation-based approach. Mol. Ecol. Resour. 20, 154–169 (2020).
    Google Scholar 
    Simillion, C., Liechti, R., Lischer, H. E. L., Ioannidis, V. & Bruggmann, R. Avoiding the pitfalls of gene set enrichment analysis with SetRank. BMC Bioinformatics 18, 151 (2017).
    Google Scholar 
    Dixon, G., Abbott, E. & Matz, M. Meta-analysis of the coral environmental stress response: Acropora corals show opposing responses depending on stress intensity. Mol. Ecol. 29, 2855–2870 (2020).
    Google Scholar 
    Rosenzweig, R., Nillegoda, N. B., Mayer, M. P. & Bukau, B. The Hsp70 chaperone network. Nat. Rev. Mol. Cell Biol. 20, 665–680 (2019).
    Google Scholar 
    Louis, Y. D. et al. Local acclimatisation-driven differential gene and protein expression patterns of Hsp70 in Acropora muricata: Implications for coral tolerance to bleaching. Mol. Ecol. 29, 4382–4394 (2020).
    Google Scholar 
    van Oppen, M. J. H. & Lough, J. M. Synthesis: Coral bleaching — Patterns, processes, causes and consequences. in coral bleaching: Patterns, processes, causes and consequences (eds. van Oppen, M. J. H. & Lough, J. M.) 175–176 (Springer berlin heidelberg, berlin, heidelberg, 2009).Matthews, J. L. et al. Optimal nutrient exchange and immune responses operate in partner specificity in the cnidarian-dinoflagellate symbiosis. Proc. Natl. Acad. Sci. USA. 114, 13194–13199 (2017).
    Google Scholar 
    Matthews, J. L. et al. Partner switching and metabolic flux in a model cnidarian–dinoflagellate symbiosis. Proceedings of the Royal Society B: Biological Sciences 285, 20182336 (2018).
    Google Scholar 
    Hillyer, K. E., Dias, D., Lutz, A., Roessner, U. & Davy, S. K. 13C metabolomics reveals widespread change in carbon fate during coral bleaching. Metabolomics 14, 12 (2017).
    Google Scholar 
    Hillyer, K. E. et al. Metabolite profiling of symbiont and host during thermal stress and bleaching in the coral Acropora aspera. Coral Reefs 36, 105–118 (2017).
    Google Scholar 
    González-Pech, R. A. et al. Physiological factors facilitating the persistence of Pocillopora aliciae and Plesiastrea versipora in temperate reefs of south-eastern australia under ocean warming. Coral Reefs 41, 1239–1253 (2022).
    Google Scholar 
    Rädecker, N. et al. Heat stress destabilizes symbiotic nutrient cycling in corals. Proc. Natl. Acad. Sci. USA. 118, e2022653118 (2021).
    Google Scholar 
    Breiman, L. Random Forests. Mach. Learn. 45, 5–32 (2001).
    Google Scholar 
    Pinsky, M. L., Clark, R. D. & Bos, J. T. Coral Reef Population Genomics in an Age of Global Change. Annu. Rev. Genet. 57, 87–115 (2023).
    Google Scholar 
    Loya, Y. et al. Coral bleaching: the winners and the losers. Ecol. Lett. 4, 122–131 (2001).
    Google Scholar 
    van Woesik, R., Sakai, K., Ganase, A. & Loya, Y. Revisiting the winners and the losers a decade after coral bleaching. Mar. Ecol. Prog. Ser. 434, 67–76 (2011).
    Google Scholar 
    McClanahan, T. R. et al. Highly variable taxa-specific coral bleaching responses to thermal stresses. Mar. Ecol. Prog. Ser. 648, 135–151 (2020).
    Google Scholar 
    McClanahan, T. R. et al. Large geographic variability in the resistance of corals to thermal stress. Glob. Ecol. Biogeogr. 29, 2229–2247 (2020).
    Google Scholar 
    González-Rivero, M. et al. The catlin seaview survey – kilometre-scale seascape assessment, and monitoring of coral reef ecosystems. Aquat. Conserv. 24, 184–198 (2014).
    Google Scholar 
    Nakagawa, S., Johnson, P. C. D. & Schielzeth, H. The coefficient of determination R2 and intra-class correlation coefficient from generalized linear mixed-effects models revisited and expanded. J. R. Soc. Interface 14, 20170213 (2017).
    Google Scholar 
    Renema, W. et al. Are coral reefs victims of their own past success?. Sci Adv 2, e1500850 (2016).
    Google Scholar 
    Mellin, C. et al. Cumulative risk of future bleaching for the world’s coral reefs. Sci Adv 10, eadn9660 (2024).
    Google Scholar 
    Williams, D., Nedimyer, K., Bright, A. & Ladd, M. Genotypic inventory and impact of the 2023 marine heatwave on Acropora palmata (elkhorn coral) populations in the Upper Florida Keys, USA: 2020-2023 (National Oceanic and Atmospheric Administration, USA, 2024) https://doi.org/10.25923/37C0-X182.Logan, C. A., Dunne, J. P., Ryan, J. S., Baskett, M. L. & Donner, S. D. Quantifying global potential for coral evolutionary response to climate change. Nat. Clim. Chang. 11, 537–542 (2021).
    Google Scholar 
    Fagan, W. F. & Holmes, E. E. Quantifying the extinction vortex. Ecol. Lett. 9, 51–60 (2006).
    Google Scholar 
    Andrello, M. et al. A global map of human pressures on tropical coral reefs. Conserv. Lett. 15, https://doi.org/10.1111/conl.12858 (2022).Claar, D. C. et al. Dynamic symbioses reveal pathways to coral survival through prolonged heatwaves. Nat. Commun. 11, 6097 (2020).
    Google Scholar 
    UNEP-WCMC & IUCN. The World Database on Protected Areas (WDPA). (2020).Beyer, H. L. et al. Risk-sensitive planning for conserving coral reefs under rapid climate change. Conserv. Lett. 11, e12587 (2018).
    Google Scholar 
    Drury, C. et al. Genomic patterns in Acropora cervicornis show extensive population structure and variable genetic diversity. Ecol. Evol. 7, 6188–6200 (2017).
    Google Scholar 
    Leinonen, R., Sugawara, H., Shumway, M. & Collaboration, I. N. S. D. The sequence read archive. Nucleic Acids Res. 39, D19–D21 (2010).
    Google Scholar 
    Andrews, S. FASTQC: A quality control tool for high throughput sequence data. (2010).Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet.journal 17, 10–12 (2011).
    Google Scholar 
    Li, H. et al. The sequence alignment/map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).
    Google Scholar 
    McKenna, A. et al. The genome analysis toolkit: A mapreduce framework for analyzing next-generation dna sequencing data. Genome Res. 20, 1297–1303 (2010).
    Google Scholar 
    Brooks, M. E. et al. glmmTMB balances speed and flexibility among packages for zero-inflated generalized linear mixed modeling. R J. 9, 378–400 (2017).
    Google Scholar 
    Goudet, J. HIERFSTAT, a package for R to compute and test hierarchical F-statistics. Mol. Ecol. Notes 5, 184–186 (2005).
    Google Scholar 
    Sully, S., Burkepile, D. E., Donovan, M. K., Hodgson, G. & van Woesik, R. A global analysis of coral bleaching over the past two decades. Nat. Commun. 10, 1–5 (2019).
    Google Scholar 
    Liu, G. et al. Reef-Scale Thermal stress monitoring of coral ecosystems: New 5-km global products from noaa coral reef watch. Remote Sensing 6, 11579–11606 (2014).
    Google Scholar 
    Frichot, E. & François, O. L. E. A. An R package for landscape and ecological association studies. Methods Ecol. Evol. 6, 925–929 (2015).
    Google Scholar 
    Storey, J. D. The positive false discovery rate: A bayesian interpretation and the q-Value. Ann. Stat. 31, 2013–2035 (2003).
    Google Scholar 
    Camacho, C. et al. BLAST+: architecture and applications. BMC Bioinformatics 10, 421 (2009).
    Google Scholar 
    Bateman, A. et al. UniProt: A hub for protein information. Nucleic Acids Res. 43, D204–D212 (2015).
    Google Scholar 
    Spalding, M. D. et al. Marine ecoregions of the world: A bioregionalization of coastal and shelf areas. Bioscience 57, 573–583 (2007).
    Google Scholar 
    Barton, K. MuMIn: Multi-model inference. (2009).Mann, H. B. & Whitney, D. R. On a test of whether one of two random variables is stochastically larger than the other. Ann. Math. Stat. 18, 50–60 (1947).
    Google Scholar 
    Unep-Wcmc, WorldFish Centre, World resources institute & the nature conservancy. Global distribution of warm-water coral reefs, compiled from multiple sources including the millennium coral reef mapping project. (UN environment world conservation monitoring centre, cambridge (UK), 2021).Provoost, P., Bosch, S. & appletans, W. Robis: R client to access data from the OBIS API. (Ocean biogeographic information system. Intergovernmental oceanographic commission of UNESCO, 2017).Hadfield, J. D. MCMC Methods for multi-response generalized linear mixed models: The MCMCglmm R Package. J. Stat. Softw. 33, 1–22 (2010).
    Google Scholar 
    Pinheiro, J., bates, D., Debroy, S. & Sarkar, D. Nlme: Nonlinear mixed-effects models. (2013).Selmoni, O., Cleves, P. & Exposito-Alonso, M. Scripts and data from ‘global coral genomic vulnerability explains recent reef losses’. https://doi.org/10.5281/zenodo.10838947 (2024).
    Google Scholar 
    NOAA national centers for environmental information. ETOPO 2022 15 arc-second global relief Model. https://doi.org/10.25921/fd45-gt74 (2022).Unep-Wcmc, WorldFish-Center, Wri & Tnc. Global distribution of warm-water coral reefs, compiled from multiple sources including the millennium coral reef mapping project. Version 4.1. http://data.unep-wcmc.org/datasets/1 (2021).Download referencesAcknowledgementsWe are grateful to the openness of many researchers who make genomic data publicly available, making this research possible: Cooke et al., Drury et al., Fulle et al., Matz et al., Selmoni et al., Shinzato et al., and Torquato et al. We also thank the Catlin Seaview Survey project for collecting and giving access to the field survey data for the Acropora GBR case study, the Coral Reef Watch for giving access to the degree heating week data, the United Nations Environment Programme World Conservation Monitoring Centre for giving access to the worldwide distribution of coral reefs data, and Dixon et al. for sharing the Acropora gene expression data. We thank Rachael Bay and Stephane Joost for early discussions of coral datasets, and thank the MoiLab and Cleves lab for comments and discussions. M.E.-A. is supported by the Office of the Director of the National Institutes of Health’s Early Investigator Award (1DP5OD029506-01), the Carnegie Institution for Science, the Howard Hughes Medical Institute, and the University of California, Berkeley. Computational analyses were done on the High-Performance Computing clusters Memex, Calc, and MoiNode supported by the Carnegie Institution for Science. P.A.C. is supported by an NSF-EDGE grant (2128073), Pew Biomedical and Marine Fellowship (00036631), Revive and Restore, the Carnegie Institution for Science, and Moore Foundation grant (12187). We also thank a Carnegie Venture Grant (P.A.C. and M.E.-A.) for support.Author informationAuthor notesThese authors jointly supervised this work: Phillip A. Cleves, Moises Exposito-Alonso.Authors and AffiliationsDepartment of Plant Biology, Carnegie Institution for Science, Stanford, CA, USAOliver Selmoni & Moises Exposito-AlonsoDepartment of Embryology, Carnegie Institution for Science, Baltimore, MD, USAOliver Selmoni & Phillip A. ClevesDepartment of Biology, Johns Hopkins University, Baltimore, MD, USAOliver Selmoni & Phillip A. ClevesDepartment of Molecular and Cell Biology, University of California Berkeley, Berkeley, CA, USAOliver Selmoni & Phillip A. ClevesDepartment of Integrative Biology, University of California Berkeley, Berkeley, CA, USAOliver Selmoni & Moises Exposito-AlonsoDepartment of Biology, Stanford University, Stanford, CA, USAMoises Exposito-AlonsoDepartment of Global Ecology, Carnegie Institution for Science, Stanford, CA, USAMoises Exposito-AlonsoHoward Hughes Medical Institute, University of California Berkeley, Berkeley, CA, USAMoises Exposito-AlonsoAuthorsOliver SelmoniView author publicationsSearch author on:PubMed Google ScholarPhillip A. ClevesView author publicationsSearch author on:PubMed Google ScholarMoises Exposito-AlonsoView author publicationsSearch author on:PubMed Google ScholarContributionsO.S., P.A.C., and M.E.-A. conceived and led the project. O.S. conducted research, O.S., P.A.C., and M.E.-A. interpreted the results and wrote the manuscript.Corresponding authorsCorrespondence to
    Oliver Selmoni, Phillip A. Cleves or Moises Exposito-Alonso.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 InformationDescription of Additional Supplementary InformationSupplementary Data 1Transparent Peer Review fileRights and permissions
    Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
    Reprints and permissionsAbout this articleCite this articleSelmoni, O., Cleves, P.A. & Exposito-Alonso, M. Global coral genomic vulnerability explains recent reef losses.
    Nat Commun (2025). https://doi.org/10.1038/s41467-025-67616-5Download citationReceived: 05 March 2025Accepted: 04 December 2025Published: 19 December 2025DOI: https://doi.org/10.1038/s41467-025-67616-5Share 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

    Reconstruction of 2,965 Microbial Genomes from Mangrove Sediments across Guangxi, China

    AbstractMangrove sediments, being organic-rich and anoxic, host diverse and functionally important microorganisms that play crucial roles in global biogeochemical cycling. In order to characterize this diversity at the genome-resolved level, we collected 38 sediment samples encompassing both surface (0–5 cm) and core (up to 90 cm) depths from six representative mangrove sites across Guangxi Province, China. Using a standardized pipeline for assembly, binning, and dereplication, we reconstructed 2,965 non-redundant metagenome-assembled genomes (MAGs), comprising 2,383 bacterial and 582 archaeal genomes spanning 78 microbial phyla. This dataset captures the high microbial diversity and functional potential within mangrove sediments under variable environmental conditions. It provides a valuable genomic resource for investigating the structure, metabolism, and ecological roles of sediment microbial communities in intertidal, nutrient-rich ecosystems, supporting future studies on microbial adaptation and biogeochemical cycling in global blue carbon environments.

    Data availability

    The raw sequencing dataset has been deposited in NCBI (PRJNA1270782), and the metagenome-assembled genomes (MAGs) have been deposited in the ENA (PRJEB96880) and the figshare database (https://doi.org/10.6084/m9.figshare.29320385).
    Code availability

    All in-house code used in this paper is available through a GitHub repository at https://github.com/SongzeCHEN/MetaGenome-MAG-Analysis.
    ReferencesAlongi, D. M. Carbon Cycling and Storage in Mangrove Forests. Annu. Rev. Mar. Sci. 6, 195–219 (2014).
    Google Scholar 
    Giri, C. et al. Status and distribution of mangrove forests of the world using earth observation satellite data. Glob. Ecol. Biogeogr. 20, 154–159 (2011).
    Google Scholar 
    Palit, K., Rath, S., Chatterjee, S. & Das, S. Microbial diversity and ecological interactions of microorganisms in the mangrove ecosystem: Threats, vulnerability, and adaptations. Environ. Sci. Pollut. Res. 29, 32467–32512 (2022).
    Google Scholar 
    Zhang, Z.-F., Liu, L.-R., Pan, Y.-P., Pan, J. & Li, M. Long-read assembled metagenomic approaches improve our understanding on metabolic potentials of microbial community in mangrove sediments. Microbiome 11, 188 (2023).
    Google Scholar 
    Brander, L. M. et al. Ecosystem service values for mangroves in Southeast Asia: A meta-analysis and value transfer application. Ecosyst. Serv. 1, 62–69 (2012).
    Google Scholar 
    Mai, Z. et al. Characteristics of Microbial Community and Function With the Succession of Mangroves. Front. Microbiol. 12 (2021).Liu, Y. et al. Bacterial Community Structure and Environmental Driving Factors in the Surface Sediments of Six Mangrove Sites from Guangxi, China. Microorganisms 12, 2607 (2024).
    Google Scholar 
    Zhang, Z.-F., Pan, J., Pan, Y.-P. & Li, M. Biogeography, Assembly Patterns, Driving Factors, and Interactions of Archaeal Community in Mangrove Sediments. mSystems 6, e01381–20 (2021).
    Google Scholar 
    Zhang, Z.-F., Pan, Y.-P., Liu, Y. & Li, M. High-Level Diversity of Basal Fungal Lineages and the Control of Fungal Community Assembly by Stochastic Processes in Mangrove Sediments. Appl. Environ. Microbiol. 87, e00928–21 (2021).
    Google Scholar 
    Reis, C. R. G., Nardoto, G. B. & Oliveira, R. S. Global overview on nitrogen dynamics in mangroves and consequences of increasing nitrogen availability for these systems. Plant Soil 410, 1–19 (2017).
    Google Scholar 
    The Genome Standards Consortium. et al. Minimum information about a single amplified genome (MISAG) and a metagenome-assembled genome (MIMAG) of bacteria and archaea. Nat. Biotechnol. 35, 725–731 (2017).
    Google Scholar 
    Chen, S., Zhou, Y., Chen, Y. & Gu, J. fastp: an ultra-fast all-in-one FASTQ preprocessor. Bioinformatics 34, i884–i890 (2018).
    Google Scholar 
    Li, D., Liu, C.-M., Luo, R., Sadakane, K. & Lam, T.-W. MEGAHIT: an ultra-fast single-node solution for large and complex metagenomics assembly via succinct de Bruijn graph. Bioinformatics 31, 1674–1676 (2015).
    Google Scholar 
    Shen, W., Le, S., Li, Y. & Hu, F. SeqKit: A Cross-Platform and Ultrafast Toolkit for FASTA/Q File Manipulation. PLOS ONE 11, e0163962 (2016).
    Google Scholar 
    Gurevich, A., Saveliev, V., Vyahhi, N. & Tesler, G. QUAST: quality assessment tool for genome assemblies. Bioinformatics 29, 1072–1075 (2013).
    Google Scholar 
    Wu, Y.-W., Simmons, B. A. & Singer, S. W. MaxBin 2.0: an automated binning algorithm to recover genomes from multiple metagenomic datasets. Bioinformatics 32, 605–607 (2016).
    Google Scholar 
    Kang, D. D. et al. MetaBAT 2: an adaptive binning algorithm for robust and efficient genome reconstruction from metagenome assemblies. PeerJ 7, e7359 (2019).
    Google Scholar 
    Alneberg, J. et al. Binning metagenomic contigs by coverage and composition. Nat. Methods 11, 1144–1146 (2014).
    Google Scholar 
    Pan, S., Zhu, C., Zhao, X.-M. & Coelho, L. P. A deep siamese neural network improves metagenome-assembled genomes in microbiome datasets across different environments. Nat. Commun. 13, 2326 (2022).
    Google Scholar 
    Qiu, Z. et al. BASALT refines binning from metagenomic data and increases resolution of genome-resolved metagenomic analysis. Nat. Commun. 15, 2179 (2024).
    Google Scholar 
    Chklovski, A., Parks, D. H., Woodcroft, B. J. & Tyson, G. W. CheckM2: a rapid, scalable and accurate tool for assessing microbial genome quality using machine learning. Nat. Methods 20, 1203–1212 (2023).
    Google Scholar 
    Aroney, S. T. N. et al. CoverM: read alignment statistics for metagenomics. Bioinformatics 41, btaf147 (2025).
    Google Scholar 
    Nissen, J. N. et al. Improved metagenome binning and assembly using deep variational autoencoders. Nat. Biotechnol. 39, 555–560 (2021).
    Google Scholar 
    Sieber, C. M. K. et al. Recovery of genomes from metagenomes via a dereplication, aggregation and scoring strategy. Nat. Microbiol. 3, 836–843 (2018).
    Google Scholar 
    Liu, Y. et al. Expanded diversity of Asgard archaea and their relationships with eukaryotes. Nature 593, 553–557 (2021).
    Google Scholar 
    Vollmers, J., Wiegand, S., Lenk, F. & Kaster, A.-K. How clear is our current view on microbial dark matter? (Re-)assessing public MAG & SAG datasets with MDMcleaner. Nucleic Acids Res. 50, e76–e76 (2022).
    Google Scholar 
    Olm, M. R., Brown, C. T., Brooks, B. & Banfield, J. F. dRep: a tool for fast and accurate genomic comparisons that enables improved genome recovery from metagenomes through de-replication. ISME J. 11, 2864–2868 (2017).
    Google Scholar 
    Albanese, D. & Donati, C. Large-scale quality assessment of prokaryotic genomes with metashot/prok-quality. F1000Research 10, 822 (2021).
    Google Scholar 
    Orakov, A. et al. GUNC: detection of chimerism and contamination in prokaryotic genomes. Genome Biol. 22, 178 (2021).
    Google Scholar 
    Lowe, T. M. & Chan, P. P. tRNAscan-SE On-line: integrating search and context for analysis of transfer RNA genes. Nucleic Acids Res. 44, W54–W57 (2016).
    Google Scholar 
    Rinke, C. et al. A standardized archaeal taxonomy for the Genome Taxonomy Database. Nat. Microbiol. 6, 946–959 (2021).
    Google Scholar 
    Vasimuddin, M., Misra, S., Li, H. & Aluru, S. Efficient Architecture-Aware Acceleration of BWA-MEM for Multicore Systems. in 2019 IEEE International Parallel and Distributed Processing Symposium (IPDPS) 314–324, https://doi.org/10.1109/IPDPS.2019.00041 (IEEE, Rio de Janeiro, Brazil, 2019).Hyatt, D. et al. Prodigal: prokaryotic gene recognition and translation initiation site identification. BMC Bioinformatics 11, 119 (2010).
    Google Scholar 
    Sunagawa, S. et al. Metagenomic species profiling using universal phylogenetic marker genes. Nat. Methods 10, 1196–1199 (2013).
    Google Scholar 
    Mende, D. R., Sunagawa, S., Zeller, G. & Bork, P. Accurate and universal delineation of prokaryotic species. Nat. Methods 10, 881–884 (2013).
    Google Scholar 
    Martinez-Gutierrez, C. A. & Aylward, F. O. Phylogenetic Signal, Congruence, and Uncertainty across Bacteria and Archaea. Mol. Biol. Evol. 38, 5514–5527 (2021).
    Google Scholar 
    Capella-Gutiérrez, S., Silla-Martínez, J. M. & Gabaldón, T. trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25, 1972–1973 (2009).
    Google Scholar 
    Wong, T. K. F. et al. IQ-TREE 3: Phylogenomic Inference Software using Complex Evolutionary Models (2025).Letunic, I. & Bork, P. Interactive Tree of Life (iTOL) v6: recent updates to the phylogenetic tree display and annotation tool. Nucleic Acids Res. 52, W78–W82 (2024).
    Google Scholar 
    NCBI BioProject https://identifiers.org/ncbi/bioproject:PRJNA1270782 (2025).NCBI Sequence Read Archive https://identifiers.org/ncbi/insdc.sra:SRP589204 (2017).Liu, Y. et al. Reconstruction of 2,965 Microbial Genomes from Mangrove Sediments across Guangxi, China. figshare https://doi.org/10.6084/m9.figshare.29320385 (2025).ENA European Nucleotide Archive https://identifiers.org/ena.embl:PRJEB96880 (2025).Download referencesAcknowledgementsThis work was supported by the Natural Science Foundation of Guangxi, China (Project Nos. 2024GXNSFBA010371 and 2025GXNSFHA069226), the Beibu Gulf University High-level Talent Scientific Research Start-up Project (Project No. 23KYQD18), the Natural Science Foundation of Guangxi, China (Project No. 2025GXNSFHA069232), the Improving the Basic Scientific Research Capability of Young and Middle-aged Teachers in Guangxi Colleges Project (Project No. 2025KY0471), Science and Technology Bases and Talents Special Project in Guangxi (AD22035181), Marine Science and Technology Innovation Cooperation Foundation of Beibu Gulf Project (Project No. 03190010), Development of Utilization Technology of Probiotics in Fish Gut of Symbiotic System Project (Project No. 02040772), and National College Student Innovation and Entrepreneurship Training Program (Project No. S202511607100 and S202411607011).Author informationAuthor notesThese authors contributed equally: Ying Liu, Songze Chen, Huiquan Li.Authors and AffiliationsPinglu Canal and Beibu Gulf Coastal Ecosystem Observation and Research Station of Guangxi, Guangxi Key Laboratory of Marine Environmental Disaster Processes and Ecological Protection Technology, College of Marine Sciences, Beibu Gulf University, Qinzhou, 535011, ChinaYing Liu, Yue Sun, Yunru Li, Jingjing Song, Dan Sun, Mingzhong Liang, Jianqing Chen, Bin Gong & Rongping BuShenzhen Ecological and Environmental Monitoring Center of Guangdong Province, Shenzhen, 518049, ChinaSongze ChenShenzhen Key Laboratory of Marine Archaea Geo-Omics, Department of Ocean Science and Department of Ocean Science & Engineering, Southern University of Science and Technology, Shenzhen, 518055, ChinaHuiquan Li & Nazia MahtabFangchenggang Vocational and Technical College, Fangchenggang, 538000, ChinaJing SunStomatological Center, Peking University Shenzhen Hospital, Shenzhen, 518036, ChinaJiaojiao JingAuthorsYing LiuView author publicationsSearch author on:PubMed Google ScholarSongze ChenView author publicationsSearch author on:PubMed Google ScholarHuiquan LiView author publicationsSearch author on:PubMed Google ScholarNazia MahtabView author publicationsSearch author on:PubMed Google ScholarYue SunView author publicationsSearch author on:PubMed Google ScholarYunru LiView author publicationsSearch author on:PubMed Google ScholarJingjing SongView author publicationsSearch author on:PubMed Google ScholarDan SunView author publicationsSearch author on:PubMed Google ScholarMingzhong LiangView author publicationsSearch author on:PubMed Google ScholarJianqing ChenView author publicationsSearch author on:PubMed Google ScholarJing SunView author publicationsSearch author on:PubMed Google ScholarBin GongView author publicationsSearch author on:PubMed Google ScholarJiaojiao JingView author publicationsSearch author on:PubMed Google ScholarRongping BuView author publicationsSearch author on:PubMed Google ScholarContributionsY.L., S.C. and R.B. designed the study. Y.L., S.C., H.L., Y.S., J.J. and R.B. performed the data analysis, prepared the figure and tables, wrote the paper, and revised the manuscript. Y.L. and S.C. collected the samples and conducted the experimental procedures. S.C., H.L., N.M., Y.L., J.S., D.S., M.L., J.C., J.S., B.G., J.J. and R.B. provided helpful comments and suggestions to improve the manuscript. All co-authors read and approved the final manuscript.Corresponding authorsCorrespondence to
    Jiaojiao Jing or Rongping Bu.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Additional informationPublisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary informationTable S1Table S2Table S3Table S4Rights 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 articleLiu, Y., Chen, S., Li, H. et al. Reconstruction of 2,965 Microbial Genomes from Mangrove Sediments across Guangxi, China.
    Sci Data (2025). https://doi.org/10.1038/s41597-025-06438-yDownload citationReceived: 19 June 2025Accepted: 10 December 2025Published: 19 December 2025DOI: https://doi.org/10.1038/s41597-025-06438-yShare 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

    Unexpected microbial rhodopsin dynamics in sync with phytoplankton blooms

    AbstractThe surface ocean is the largest sunlit environment on Earth where marine microalgae are known as the main drivers of global productivity. However, rhodopsin phototrophs are actually the most abundant metabolic group, suggesting a major role in the biogeochemical cycles. While previous studies have shown that rhodopsin-containing bacterioplankton thrive in the most severely nutrient-depleted environments, growing evidence suggest that this type of phototrophy may also be relevant in nutrient-rich environments. To examine its role in productive waters, we investigated the monthly rhodopsin dynamics in the upwelling system of the Southern California Bight by measuring retinal–the photoreactive chromophore essential for rhodopsin function–in seawater. Unlike oligotrophic regions, rhodopsin levels peaked during the highly productive spring phytoplankton bloom, coinciding with the highest chlorophyll concentrations. Heterotrophic bacterial abundances, particularly within the order Flavobacteriales, correlated strongly with rhodopsin concentrations, allowing us to build linear models to predict rhodopsin distributions in a productive environment. Metagenomic data further showed that Flavobacteriales also dominated the rhodopsin gene pool when the highest rhodopsin levels were recorded, underscoring their key contribution to light-driven energy capture. Overall, our findings reveal that rhodopsin phototrophy plays a substantial role in productive marine systems, broadening its recognized importance far beyond oligotrophic oceans.

    Data availability

    Source data are provided with this paper. 16S rDNA amplicon and shotgun sequencing data are available on Genbank (https://www.ncbi.nlm.nih.gov/genbank/) under the Bioproject PRJNA1040444. Metagenome Assembled Genomes (MAGs) are available on Figshare https://doi.org/10.6084/m9.figshare.2985686691 Source data are provided with this paper.
    ReferencesFalkowski, P. G. The role of phytoplankton photosynthesis in global biogeochemical cycles. Photosynth Res. 39, 235–258 (1994).
    Google Scholar 
    Larkum, A., Ritchie, R. & Raven, J. J. P. Living off the sun: chlorophylls, bacteriochlorophylls and rhodopsins. Photosynthetica 56, 11–43 (2018).
    Google Scholar 
    Karl, D. M. Solar energy capture and transformation in the sea. Elementa 2, 000021 (2014).
    Google Scholar 
    Béjà, O. et al. Bacterial rhodopsin: evidence for a new type of phototrophy in the sea. Science 289, 1902–1906 (2000).
    Google Scholar 
    Martinez, A., Bradley, A. S., Waldbauer, J. R., Summons, R. E. & DeLong, E. F. Proteorhodopsin photosystem gene expression enables photophosphorylation in a heterologous host. Proc. Natl. Acad. Sci. USA 104, 5590–5595 (2007).
    Google Scholar 
    Finkel, O. M., Béjà, O. & Belkin, S. Global abundance of microbial rhodopsins. ISME J. 7, 448–451 (2013).
    Google Scholar 
    Kandori, H. Ion-pumping microbial rhodopsins. Front. Mol. Biosci. 2, 52 (2015).
    Google Scholar 
    Pinhassi, J., DeLong, E. F., Béjà, O., Gonzalez, J. M. & Pedrós-Alió, C. Marine bacterial and archaeal ion-pumping rhodopsins: genetic diversity, physiology, and ecology. Microbiol. Mol. Biol. Rev. 80, 929–954 (2016).
    Google Scholar 
    Gómez-Consarnau, L. et al. Microbial rhodopsins are major contributors to the solar energy captured in the sea. Sci. Adv. 5, eaaw8855 (2019).
    Google Scholar 
    Kirchman, D. L. & Hanson, T. E. Bioenergetics of photoheterotrophic bacteria in the oceans. Environ. Microbiol. Rep. 5, 188–199 (2013).
    Google Scholar 
    Morris, R. M. et al. Comparative metaproteomics reveals ocean-scale shifts in microbial nutrient utilization and energy transduction. ISME J. 4, 673–685 (2010).
    Google Scholar 
    Martínez-García, S. & Pinhassi, J. Adaptations of microorganisms to low nutrient environments: managing life in the oligotrophic ocean. In Encyclopedia of Microbiology 4th edn, (ed. Schmidt, T. M.) (Academic Press, Oxford, 2019).Brindefalk, B. et al. Distribution and expression of microbial rhodopsins in the Baltic Sea and adjacent waters. Environ. Microbiol. 18, 4442–4455 (2016).
    Google Scholar 
    Nguyen, D. et al. Winter diversity and expression of proteorhodopsin genes in a polar ocean. ISME J. 9, 1835–1845 (2015).
    Google Scholar 
    Campbell, B. J., Waidner, L. A., Cottrell, M. T. & Kirchman, D. L. Abundant proteorhodopsin genes in the North Atlantic Ocean. Environ. Microbiol. 10, 99–109 (2008).
    Google Scholar 
    Dubinsky, V. et al. Metagenomic analysis reveals unusually high incidence of proteorhodopsin genes in the ultraoligotrophic Eastern Mediterranean Sea. Environ. Microbiol. 19, 1077–1090 (2017).
    Google Scholar 
    Lami, R., Cottrell, M. T., Campbell, B. J. & Kirchman, D. L. Light-dependent growth and proteorhodopsin expression by Flavobacteria and SAR11 in experiments with Delaware coastal waters. Environ. Microbiol. 11, 3201–3209 (2009).
    Google Scholar 
    Teeling, H. et al. Substrate-controlled succession of marine bacterioplankton populations induced by a phytoplankton bloom. Science 336, 608–611 (2012).
    Google Scholar 
    Steindler, L., Schwalbach, M. S., Smith, D. P., Chan, F. & Giovannoni, S. J. Energy starved Candidatus Pelagibacter ubique substitutes light-mediated ATP production for endogenous carbon respiration. PLoS ONE 6, e19725 (2011).
    Google Scholar 
    Sieradzki, E. T., Fuhrman, J. A., Rivero-Calle, S. & Gómez-Consarnau, L. Proteorhodopsins dominate the expression of phototrophic mechanisms in seasonal and dynamic marine picoplankton communities. PeerJ 6, e5798 (2018).
    Google Scholar 
    Giovannoni, S. J. SAR11 bacteria: the most abundant plankton in the oceans. Ann. Rev. Mar. Sci. 9, 231–255 (2017).
    Google Scholar 
    Gómez-Consarnau, L. et al. Proteorhodopsin light-enhanced growth linked to vitamin-B1 acquisition in marine Flavobacteria. ISME J. 10, 1102–1112 (2016).
    Google Scholar 
    Gómez-Consarnau, L. et al. Light stimulates growth of proteorhodopsin-containing marine Flavobacteria. Nature 445, 210–213 (2007).
    Google Scholar 
    Hassanzadeh, B. et al. Microbial rhodopsins are increasingly favoured over chlorophyll in high nutrient low chlorophyll waters. Environ. Microbiol. Rep. 13, 401–406 (2021).
    Google Scholar 
    Andrew, S. M. et al. Widespread use of proton-pumping rhodopsin in Antarctic phytoplankton. Proc. Natl. Acad. Sci. USA 120, e2307638120 (2023).
    Google Scholar 
    Longhurst, A., Sathyendranath, S., Platt, T. & Caverhill, C. An estimate of global primary production in the ocean from satellite radiometer data. J. Plankton Res. 17, 1245–1271 (1995).
    Google Scholar 
    Capone, D. G. & Hutchins, D. A. Microbial biogeochemistry of coastal upwelling regimes in a changing ocean. Nat. Geosci. 6, 711–717 (2013).
    Google Scholar 
    Chow, C.-E. T. et al. Temporal variability and coherence of euphotic zone bacterial communities over a decade in the Southern California Bight. ISME J. 7, 2259 (2013).
    Google Scholar 
    Cram, J. A. et al. Seasonal and interannual variability of the marine bacterioplankton community throughout the water column over ten years. ISME J. 9, 563 (2015).
    Google Scholar 
    Countway, P. D., Vigil, P. D., Schnetzer, A., Moorthi, S. D. & Caron, D. A. Seasonal analysis of protistan community structure and diversity at the USC Microbial Observatory (San Pedro Channel, North Pacific Ocean). Limnol. Oceanogr. 55, 2381–2396 (2010).
    Google Scholar 
    Maresca, J. A., Miller, K. J., Keffer, J. L., Sabanayagam, C. R. & Campbell, B. J. Distribution and diversity of rhodopsin-producing microbes in the Chesapeake Bay. Appl. Environ. Microbiol. 84, 00137–00118 (2018).
    Google Scholar 
    Gómez-Consarnau, L., Needham, D. M., Weber, P. K., Fuhrman, J. A. & Mayali, X. Influence of Light on particulate organic matter utilization by attached and free-living marine bacteria. Front. Microbiol. 10, 1204 (2019).
    Google Scholar 
    Strauss, J. et al. Plastid-localized xanthorhodopsin increases diatom biomass and ecosystem productivity in iron-limited surface oceans. Nat. Microbiol. 8, 2050–2066 (2023).
    Google Scholar 
    Bar-Shalom, R. et al. Rhodopsin-mediated nutrient uptake by cultivated photoheterotrophic Verrucomicrobiota. ISME J. 17, 1063–1074 (2023).
    Google Scholar 
    Bergauer, K. et al. Organic matter processing by microbial communities throughout the Atlantic water column as revealed by metaproteomics. Proc. Natl. Acad. Sci. USA 115, E400–E408 (2018).
    Google Scholar 
    Arístegui, J. et al. Variability in plankton community structure, metabolism, and vertical carbon fluxes along an upwelling filament (Cape Juby, NW Africa). Prog. Oceanogr. 62, 95–114 (2004).
    Google Scholar 
    Giovannoni, S. J. et al. Proteorhodopsin in the ubiquitous marine bacterium SAR11. Nature 438, 82–85 (2005).
    Google Scholar 
    Béjà, O., Spudich, E. N., Spudich, J. L., Leclerc, M. & DeLong, E. F. Proteorhodopsin phototrophy in the ocean. Nature 411, 786–789 (2001).
    Google Scholar 
    Lauro, F. M. et al. The genomic basis of trophic strategy in marine bacteria. Proc. Natl. Acad. Sci. USA 106, 15527–15533 (2009).
    Google Scholar 
    Olson, D. K. et al. Proteorhodopsin variability and distribution in the North Pacific Subtropical Gyre. ISME J. 12, 1047–1060 (2018).
    Google Scholar 
    Rozenberg, A., Inoue, K., Kandori, H. & Béjà, O. Microbial rhodopsins: the last two decades. Annu. Rev. Microbiol. 75, 427–447 (2021).
    Google Scholar 
    Mannen, K. et al. Multiple roles of a conserved glutamate residue for unique biophysical properties in a new group of microbial rhodopsins homologous to TAT rhodopsin. J. Mol. Biol. 436, 168331 (2024).
    Google Scholar 
    Kolber, Z. Energy cycle in the ocean: powering the microbial world. Oceanography 20, 79–88 (2007).
    Google Scholar 
    Vader, A., Laughinghouse, H. D., Griffiths, C., Jakobsen, K. S. & Gabrielsen, T. M. Proton-pumping rhodopsins are abundantly expressed by microbial eukaryotes in a high-Arctic fjord. Environ. Microbiol. 20, 890–902 (2018).
    Google Scholar 
    Marchetti, A. et al. Marine diatom proteorhodopsins and their potential role in coping with low iron availability. ISME J. 9, 2745–2748 (2015).
    Google Scholar 
    Fernández-Gomez, B. et al. Ecology of marine Bacteroidetes: a comparative genomics approach. ISME J. 7, 1026–1037 (2013).
    Google Scholar 
    Arandia‐Gorostidi, N. et al. Light supports cell‐integrity and growth rates of taxonomically diverse coastal photoheterotrophs. Environ. Microbiol. 22, 3823–3837 (2020).González, J. M. et al. Genome analysis of the proteorhodopsin-containing marine bacterium Polaribacter sp. MED152 (Flavobacteria). Proc. Natl. Acad. Sci. USA 105, 8724–8729 (2008).
    Google Scholar 
    Mary, I. et al. Light enhanced amino acid uptake by dominant bacterioplankton groups in surface waters of the Atlantic Ocean. FEMS Microbiol. Ecol. 63, 36–45 (2008).
    Google Scholar 
    Gómez-Pereira, P. R. et al. Comparable light stimulation of organic nutrient uptake by SAR11 and Prochlorococcus in the North Atlantic subtropical gyre. ISME J. 7, 603–614 (2013).
    Google Scholar 
    Knap, A., Michaels, A., Close, A., Ducklow, H. & Dickson, A. Protocols for the Joint Global Ocean Flux Study (JGOFS) Core Measurements. IOC Manuals and Guides No. 29, UNESCO, Paris (1994).Schlitzer, R. Ocean Data View. https://odv.awi.de (2018).Strickland, J. D. & Parsons, T. R. A Practical Handbook of Seawater Analysis. 2nd ed. Bulletin No. 167. (Ottawa: Fisheries Research Board of Canada, 1972).Cruaud, P., Rasplus, J.-Y., Rodriguez, L. J. & Cruaud, A. High-throughput sequencing of multiple amplicons for barcoding and integrative taxonomy. Sci. Rep. 7, 1–12 (2017).
    Google Scholar 
    Parada, A. E., Needham, D. M. & Fuhrman, J. A. Every base matters: assessing small subunit rRNA primers for marine microbiomes with mock communities, time series and global field samples. Environ. Microbiol. 18, 1403–1414 (2016).
    Google Scholar 
    Andrews, S. FastQC: a quality control tool for high throughput sequence data. http://www.bioinformatics.babraham.ac.uk/projects/fastqc (2010).Bolger, A. M., Lohse, M. & Usadel, B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120 (2014).
    Google Scholar 
    Li, D. et al. MEGAHIT v1. 0: a fast and scalable metagenome assembler driven by advanced methodologies and community practices. Methods 102, 3–11 (2016).
    Google Scholar 
    Arkin, A. P. et al. KBase: the United States department of energy systems biology knowledgebase. Nat. Biotechnol. 36, 566–569 (2018).
    Google Scholar 
    West, P. T., Probst, A. J., Grigoriev, I. V., Thomas, B. C. & Banfield, J. F. Genome-reconstruction for eukaryotes from complex natural microbial communities. Genome Res. 28, 569–580 (2018).
    Google Scholar 
    Zhu, W., Lomsadze, A. & Borodovsky, M. Ab initio gene identification in metagenomic sequences. Nucleic Acids Res. 38, e132 (2010).
    Google Scholar 
    Hyatt, D. et al. Prodigal: prokaryotic gene recognition and translation initiation site identification. BMC Bioinforma. 11, 1–11 (2010).
    Google Scholar 
    Mirdita, M., Steinegger, M., Breitwieser, F., Söding, J. & Levy Karin, E. Fast and sensitive taxonomic assignment to metagenomic contigs. Bioinformatics 37, 3029–3031 (2021).
    Google Scholar 
    Li, W. et al. The EMBL-EBI bioinformatics web and programmatic tools framework. Nucleic Acids Res. 43, W580–W584 (2015).
    Google Scholar 
    Buchfink, B., Xie, C. & Huson, D. H. Fast and sensitive protein alignment using DIAMOND. Nat. Methods 12, 59–60 (2015).
    Google Scholar 
    Suzek, B. E. et al. UniRef clusters: a comprehensive and scalable alternative for improving sequence similarity searches. Bioinformatics 31, 926–932 (2015).
    Google Scholar 
    Levy Karin, E., Mirdita, M. & Söding, J. MetaEuk—sensitive, high-throughput gene discovery, and annotation for large-scale eukaryotic metagenomics. Microbiome 8, 1–15 (2020).
    Google Scholar 
    Groussman, R. D., Blaskowski, S., Coesel, S. N. & Armbrust, E. V. MarFERReT, an open-source, version-controlled reference library of marine microbial eukaryote functional genes. Sci. Data 10, 926 (2023).
    Google Scholar 
    Chaumeil, P. A., Mussig, A. J., Hugenholtz, P. & Parks, D. H. GTDB-Tk: a toolkit to classify genomes with the genome taxonomy database. Bioinformatics 6, 1925–1927 (2020).
    Google Scholar 
    Vasimuddin, M., Misra, S., Li, H. & Aluru, S. Efficient architecture-aware acceleration of BWA-MEM for multicore systems. In 2019 IEEE International Parallel and Distributed Processing Symposium (IPDPS) 314–324 (IEEE, Rio de Janeiro, Brazil, 2019). https://doi.org/10.1109/IPDPS.2019.00041.Li, H. The sequence alignment/map (SAM) format and SAMtools 1000 Genome Project data processing subgroup. Bioinformatics 25, 1 (2009).
    Google Scholar 
    Anders, S., Pyl, P. T. & Huber, W. HTSeq—a Python framework to work with high-throughput sequencing data. Bioinformatics 31, 166–169 (2015).
    Google Scholar 
    Ernst, O. P. et al. Microbial and animal rhodopsins: structures, functions, and molecular mechanisms. Chem. Rev. 8, 126–163 (2014).
    Google Scholar 
    Nagata, T. & Inoue, K. Rhodopsins at a glance. J. Cell Sci. 134, jcs258989 (2021).
    Google Scholar 
    Yamauchi, Y. et al. Molecular properties of a DTD channelrhodopsin from Guillardia theta. Biophys. Physicobiol. 14, 57–66 (2017).
    Google Scholar 
    Bulzu, P., Kavagutti, V. S., Andrei, A. & Ghai, R. The evolutionary kaleidoscope of rhodopsins. mSystems 7, e00405–22 (2022).
    Google Scholar 
    Needham et al. A distinct lineage of giant viruses brings a rhodopsin photosystem to unicellular marine predators. Proc. Natl. Acad. Sci. USA 116, 20574–20583 (2019).
    Google Scholar 
    Man, D. et al. Diversification and spectral tuning in marine proteorhodopsins. EMBO J. 15, 725–731 (2003).
    Google Scholar 
    Katoh, K., Misawa, K., Kuma, K. & Miyata, T. MAFFT: a novel method for rapid multiple sequence alignment based on fast Fourier transform. Nucleic Acids Res. 30, 3059–3066 (2002).
    Google Scholar 
    Castresana, J. Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Mol. Biol. Evol. 17, 540–552 (2000).
    Google Scholar 
    Johnson, L. S., Eddy, S. R. & Portugaly, E. Hidden Markov model speed heuristic and iterative HMM search procedure. BMC Bioinf. 11, 431 (2010).
    Google Scholar 
    Stamatakis, A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30, 1312–1313 (2014).
    Google Scholar 
    Kang, D. D. et al. MetaBAT 2: an adaptive binning algorithm for robust and efficient genome reconstruction from metagenome assemblies. PeerJ 7, e7359 (2019).
    Google Scholar 
    Alneberg, J. et al. Binning metagenomic contigs by coverage and composition. Nat. Methods 11, 1144–1146 (2014).
    Google Scholar 
    Pan, S., Zhao, X.-M. & Coelho, L. P. SemiBin2: self-supervised contrastive learning leads to better MAGs via deep learning. Bioinformatics 39, i21–i29 (2023).
    Google Scholar 
    Uritskiy, G. V., DiRuggiero, J. & Taylor, J. MetaWRAP-a flexible pipeline for genome-resolved metagenomic data analysis. Microbiome 6, 158 (2018).
    Google Scholar 
    Olm, M. R., Brown, C. T., Brooks, B. & Banfield, J. F. dRep: a tool for fast and accurate genomic comparisons that enables improved genome recovery from metagenomes through de-replication. ISME J. 11, 2864–2868 (2017).
    Google Scholar 
    Seemann, T. Prokka: rapid prokaryotic genome annotation. Bioinformatics 30, 2068–2069 (2014).
    Google Scholar 
    Cantalapiedra, C. P., Hernandez-Plaza, A., Letunic, I., Bork, P. & Huerta-Cepas, J. eggNOG-mapper v2: functional annotation, orthology assignments, and domain prediction at the metagenomic scale. Mol. Bio. Evol. 38, 5825–5829 (2021).
    Google Scholar 
    Aroney, S. T. N. et al. CoverM: read alignment statistics for metagenomics. Bioinformatics 41, btaf147 (2025).
    Google Scholar 
    Cuevas-Cruz, M. et al. Annotated metagenome-assembled genomes (MAGs) from the Southern California Bight. Data sets. Figshare https://doi.org/10.6084/m9.figshare.29856866.Download referencesAcknowledgementsWe thank Yamne Ortega Saad, Hiram Zayola and Lidia Montiel for their assistance in DNA extractions, statistical and bioinformatics analyses, and the crew on board the R/V Yellowfin for sample collection. This project was partially funded by the United States National Science Foundation grant (NSF, OCE1924464 and OCE-2220546), the Ministry of Economy and Competitiveness – Spanish Agencia Estatal de Investigación PID2023-152792NB-I00 and the United States-Israel Binational Science Foundation (BSF, No. 2019612).Author informationAuthors and AffiliationsMediterranean Institute for Advanced Studies, IMEDEA (UIB-CSIC), Esporles, SpainLaura Gómez-Consarnau & Estefany VillarrealMarine and Environmental Biology, Department of Biological Sciences, University of Southern California, Los Angeles, CA, USALaura Gómez-Consarnau, Babak Hassanzadeh & Sergio A. Sañudo-WilhelmyDepartamento de Oceanografía Biológica, Centro de Investigación Científica y de Educación Superior de Ensenada (CICESE), Ensenada, BC, MexicoMiguel Cuevas-CruzInstituto de Oceanografía y Cambio Global, IOCAG, Universidad de Las Palmas de Gran Canaria, Las Palmas de Gran Canaria, SpainJavier ArísteguiInstitut de Ciències del Mar (ICM), CSIC, Barcelona, SpainRamiro Logares & Francisco LatorreDepartamento de Innovación Biomédica, Centro de Investigación Científica y de Educación Superior de Ensenada (CICESE), Ensenada, Baja California, MexicoAsunción Lago-LestónDepartment of Marine Biology, Leon H. Charney School of Marine Sciences, University of Haifa, Haifa, IsraelLaura SteindlerDepartment of Earth Sciences, University of Southern California, Los Angeles, CA, USASergio A. Sañudo-WilhelmyAuthorsLaura Gómez-ConsarnauView author publicationsSearch author on:PubMed Google ScholarBabak HassanzadehView author publicationsSearch author on:PubMed Google ScholarEstefany VillarrealView author publicationsSearch author on:PubMed Google ScholarMiguel Cuevas-CruzView author publicationsSearch author on:PubMed Google ScholarJavier ArísteguiView author publicationsSearch author on:PubMed Google ScholarRamiro LogaresView author publicationsSearch author on:PubMed Google ScholarFrancisco LatorreView author publicationsSearch author on:PubMed Google ScholarAsunción Lago-LestónView author publicationsSearch author on:PubMed Google ScholarLaura SteindlerView author publicationsSearch author on:PubMed Google ScholarSergio A. Sañudo-WilhelmyView author publicationsSearch author on:PubMed Google ScholarContributionsL.G.-C., B.H., and S.A.S.-W. designed research, L.G.-C., B.H., and S.A.S.-W. collected and processed samples, L.G.-C., B.H., E.V., M.C.-C., J.A., R.L., F.L., A.L.-L., L.S. and S.A.S.-W. performed research and analyzed data; L.G.-C., B.H., and S.A.S.-W. wrote the paper with input from the other coauthors.Corresponding authorCorrespondence to
    Laura Gómez-Consarnau.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Peer review

    Peer review information
    Nature Communications thanks Shiqiang Gao, and the other, anonymous, reviewer 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 InformationDescription of Additional Supplementary FilesSupplementary Data 1Supplementary Data 2Supplementary Data 3Supplementary Data 4Supplementary Data 5Reporting SummaryTransparent Peer Review fileSource dataSource dataRights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleGómez-Consarnau, L., Hassanzadeh, B., Villarreal, E. et al. Unexpected microbial rhodopsin dynamics in sync with phytoplankton blooms.
    Nat Commun (2025). https://doi.org/10.1038/s41467-025-67474-1Download citationReceived: 03 June 2025Accepted: 01 December 2025Published: 19 December 2025DOI: https://doi.org/10.1038/s41467-025-67474-1Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Synergistic role of Trichormus variabilis and zeolites in three-layer culturing system for modulating the wastewater effluent community

    AbstractDue to their nitrogen-fixing capabilities, cyanobacteria hold significant potential for wastewater bioremediation through nutrient removal and modulation of the microbial community. The current study explored these traits using the cyanobacterium Trichormus variabilis strain AICB 1382 in combination with natural zeolites to treat municipal wastewater effluent. A combination of colorimetric, gravimetric, and 16 S/18S rDNA amplicon sequencing analyses was used to evaluate nutrient removal rates, biomass yield, and microbial community structure. The zeolites-AICB 1382 pair features (i.e. gradual release of nutrients by zeolites and vertical distribution of the cyanobacterium) enabled the stratification of the culturing system into three layers with distinct morphology and microbial populations. Results showed efficient removal of nitrate (up to 91.8%), ammonium (up to 97%), and phosphate (up to 99.2%), with enhanced biomass yields in zeolite-enriched cultures. T. variabilis reduced the diversity of the prokaryotic and eukaryotic community, lowering the presence of multidrug-resistant bacteria, whereas zeolites promoted the development of AICB 1382 and increased microbial diversity. The three-layer culturing system offers a promising solution for nutrient reclamation, biomass production, and pathogen reduction, with potential for scale-up as a semi-continuous, self-sustaining method that facilitates biomass harvesting while ensuring environmental safety for agricultural reuse or discharge into urban rivers.

    IntroductionCyanobacteria can thrive in various habitats, from dry lands1 to various water sources, where they produce oxygen and regulate the nitrate/ammonium (NO3−/NH4+) : phosphate (PO43−) ratio by nitrogen (N2) fixation. N2-fixing cyanobacteria were considered as alternatives to synthetic nitrogen fertilizers2 whose usage resulted in significant environmental pollution3 by soil acidification, humus decrement4, groundwater eutrophication3, and GHG (greenhouse gas) emissions5. Their efficacy lies in the efficiency of N2-fixation that may reach up to 60 kg/ha/season of N2 using Anabaena species6 and their ability to tolerate various, even extreme conditions7. Cyanobacteria have been widely used in bioremediation8, to reduce the levels of NH4+, NO3− (to synthesize proteins e.g., phycocyanin), PO43− and to reduce the level of pathogenic bacteria9. Although recent studies have demonstrated the benefits of Trichormus cyanobacterial extracts in promoting plant growth2, the authors emphasized the need for a phosphorus-enriched growth medium for T. variabilis. To address the natural scarcity and rapid depletion of phosphorus, they proposed utilizing wastewater as a phosphorus-rich alternative source.One way to alleviate the need to supply nutrients involves culturing microalgae (cyanobacteria included) in piggery wastewater (WW)10, municipal WW11, and aquaculture WW12. Beyond WW organic load, cyanobacterial species were selected for traits that enhance bioremediation efficiency and facilitate biomass harvesting. Many cyanobacterial species (e.g. Nostoc muscorum, Anabaena subcylindrica, A. oryzae, Spirulina platensis, and Geitlerinema sp.) have been investigated for their potential use for reclamation of WW8. Different methods were used to tackle the harvesting process, either by using cyanobacteria rich in biopolymers (i.e. slime, sheath, and capsule)13 that ease the formation of aggregates or by building cyanobacteria – trophic-related bacteria consortia14,15. Self-sustained consortia efficiently remove the biological oxygen demand (BOD) from the WW treatment plants (WWTP), clean up pollutants16 and recover nutrients coupled with mitigation of CO217. This lowers the cost for aeration, which accounts for at least 50% of the energy inputs and expenses in biological treatment plants (TP)18.This study advances nutrient reclamation and cyanobacterial biomass production by developing a three-layer system combining T. variabilis AICB 1382, zeolites as substrate and the effluent from a municipal WWTP. Given that this effluent is currently discharged into an urban river, the study focused on characterizing microbial community dynamics, including the presence and reduction of pathogenic bacteria. T. variabilis (formerly Anabaena variabilis) was selected as a suitable candidate for this study owing to its tolerance to temperature fluctuations, its biofertilizer potential2, and its capacity to grow in municipal WW environments19. The three-layer system was created by including natural zeolites, which are crystalline-hydrated aluminosilicates of alkaline and earth-alkaline elements (particularly of sodium and calcium). Due to their high capacity to exchange cations20, the zeolites have been used for culturing cyanobacteria like Arthrospira21, but also for remediation purposes22. The zeolites can adsorb cells23 and can be used for EPS-producing bacteria immobilization24. The study aimed to investigate (i) the effect of T. variabilis on nutrient reclamation and its biomass productivity; (ii) the potential of T. variabilis to inhibit the growth of harmful bacteria containing multidrug resistance genes, which occur in the effluent25, with the future aim of scaling the system before discharging the effluent into the river and, ii) the self-sustaining capacity of the culturing system based on its three-layer disposal. The analyses included the nutrient (NO3−/NH4+ and PO43−) removal rate, the biomass yield, and the prokaryotic and eukaryotic community based on 16 S rDNA/18S rDNA amplicon sequencing.ResultsThree-layer culturing system – nutrient recovery and biomass productivityStrain AICB 1382 formed buoyant, filamentous clusters in BG11 medium. In EZC system, the strain formed two distinct biomass layers separated by a clear effluent phase. The upper layer resembled a thick biofilm, composed of long, overlapping cyanobacterial filaments interspersed with bacteria, as observed by light microscopy. The lower layer on the zeolite surface appeared thin, homogeneous, and displayed an intense blue colour.Nutrient recovery analysis revealed a generally higher rate for ammonium (NH₄⁺) compared to nitrate (NO₃⁻), with overall removal rate reaching 97% and 91.8%, respectively (Fig. S1A,B; Table S1). In experiments E1 and E2, treatments containing AICB 1382 (EC and EZC) consistently outperformed the zeolite-only control (EZ) in terms of nutrient recovery (Fig. S1A−C). Specifically, nitrate removal rates ranged from 17.4 to 76.4% in EZ, 80.1–88% in EC, and 67.1–91.8% in EZC. Ammonium removal followed a similar trend, ranging from 65.1 to 76.4% in EZ, 80.2–96.8% in EC, and 67.1–97% in EZC. Phosphate (PO₄³⁻) removal was lowest in the EZ treatment (38.4–62.3%), increased substantially in EC (89.8–99.2%), and remained high in EZC (90.3–93.87%) (Fig. S1A–C; Table S1). These results underscore the synergistic effect of combining the cyanobacterium with zeolite, enhancing nutrient uptake and suggesting improved wastewater remediation potential.Biomass productivity also varied depending on the effluent and treatment applied (Fig. S1D; Table S2). The highest biomass yield was consistently recorded in the system cultured with E2E effluent, reflecting the influence of effluent composition on cyanobacterial growth. Across all experimental conditions, the lowest biomass accumulation was observed in EZ, followed by EC, with the EZC treatment producing the highest yields. Overall biomass production ranged from 47.1 mg L⁻¹ day⁻¹ in EC to 156.2 mg L⁻¹ day⁻¹ in EZC, highlighting the significant contribution of both the cyanobacterium and the zeolite substrate to enhanced growth and potential for downstream biomass utilization (Fig. S1D; Table S2).Analysis of the prokaryotic communityComposite samples analysisThe analysis of the composite samples collected by mixing all the layers showed that the treatments applied were one of the factors that shaped the taxa and their abundance in the prokaryotic community. The beta diversity analysis by PCoA (Fig. S2) matched the clusters with the treatments: effluent (E), EZ, and EC/EZC. The last two were separated in the UPGMA clustering (Fig. 1) which emphasizes sample similarity without reducing dimensions like PCoA.Fig. 1UPGMA clustering of the biomass samples collected from the three experiments (E1, E2, E3) based on the OTUs abundance using the Bray-Curtis similarity matrix. Each experiment included the effluent (E), EZ, and 2 containers (A and B) of EZC. The experiments E1 and E2 also tested the EC condition.Full size imageAlpha-diversity assay strengthened and deepened this result showing differences in taxa occurrence, abundance, diversity, and dominance (Fig. S3). The number of taxa (Chao-1 index) differed between the effluents and the treatments, but their abundance and diversity (Evenness and Shannon-H indices) showed a similar pattern for the effluent (E) and the EZ condition (Fig. S4). No dominant taxa were found in these tanks, contrary to EC/EZC conditions. Comparison between the treatments pointed out a significant difference (ANOVA test) between the effluent (E) and the EZC samples based on the Shannon-H index (F(3, 10) = 17.911, p <.001) (Table S3). Levene’s test indicated that the variances were homogenous, F(3, 10) = 3.043, p =.079; thus, Tukey’s HSD Test for multiple comparisons showed that the mean value of the Shannon-H index was significantly different (p <.001, 95% C.I. = [1.550, 4.326]).Another factor that influenced the structure of the prokaryotic community was the effluent type. Within the same cluster (EZ/EZC), the samples cultured in the first two effluents clustered separately from the third (Fig. 1), suggesting a separation due to the effluent microbial load. This fact was confirmed by the OTUs analysis (Fig. 2) where E1E and E2E shared 619 OTUs from 1389 to 1555 OTUs, making them more similar than the third effluent (482 OTUs). Even though their Chao-1 index was slightly different (Fig. S3), the first two effluents had equal OTU abundance (Evenness index). The Shannon-H index indicated a greater variety of species and a fairer distribution of individuals among species without dominant taxa in E1E and E2E than in E3E.Fig. 2The number of microbial OTUs and their common cores in the E1E, E2E, and E3E effluents. Each OTU was represented by a bullet. The numbers at the periphery indicate the unique OTUs specific to each effluent, while the overlapping areas represent the OTUs shared among the effluents.Full size imageDuring culturing, the OTUs lowered in all tanks (Fig. 3A) retaining a common core for all samples (three-point and composite) (Fig. S5A-C) as follows: 16 OTUs in E1, 32 in E2, and 26 in E3 at the end of the experiments (Fig. S5A−C), regardless of culturing conditions. Thus, these OTUs were unresponsive to the conditions tested. Except for the common core or overlapping between two or more samples, each sample had a specific number of OTUs present (Fig. S5D−F).Fig. 3The species richness (A), evenness (B), dominance (C), and diversity (D) indices based on the OTU abundances from different culturing conditions: EZ, EC, and EZC. The three-point samples were collected from the top (1), middle (2), and bottom (3) layers of the tanks. E = effluent.Full size imageThree-point samples analysisBeyond the effects of culture conditions created by zeolites, strain AICB 1382, and the effluent composition, the microbial community exhibited variation among the three layers. When AICB 1382 was present (EC/EZC) the culturing system exhibited three distinctive layers. Layers 1 and 3 were significantly different from layer 2 according to the alpha diversity indices. The evenness (Fig. 3B), and the total diversity (Fig. 3D) indices were significantly larger in layer 2 relative to the evenness (ANOVA test F(9, 26) = 13.670, p <.001) and the diversity index (ANOVA test, F(9, 26) = 15.694, p <.001) of the samples collected from layers 1 and 3 (Tables S4, S5). Levene’s test showed homogenous variances in both cases, F(9, 26) = 2.498/1.070, p =.033/0.416; thus, the Post-Hoc analyses using Tukey’s HSD test showed a significant difference (p <.001) between ECZ1/ECZ3 and the rest of the samples.The Dominance_D index (Fig. 3C) revealed that the microbial communities from layers 1 and 3 were dominated by a few taxa with high relative abundance. The SIMPER (Similarity Percentage) analysis outlined T. variabilis AICB 1382 among the top taxa that accounted for the differences among the samples (Table S6) and most probably was responsible with the large dominance index registered for layers 1 and 3. This taxon contributed the most to the overall dissimilarity among growth conditions (35.74%) from the top ten OTUs shown (49.27%). The variations in the relative abundance across layers supported the clustering patterns observed in the PCoA analysis (Fig. S6) and the UPGMA dendrogram (Fig. 4) which split the top and bottom layers from the middle layer for the EC/EZC conditions. For these treatments, middle-layer samples were partitioned by effluent type, with E1/E3 distinguished from E2.Fig. 4UPGMA hierarchical clustering of the three-point samples (in different colors) based on the OTUs abundance using Bray-Curtis similarity matrix. Each experiment included the effluent (E), EZ, and 2 containers (A and B) of EZC. The experiments E1 and E2 included the EC condition.Full size imagePhylum-level analysisPhylum-level analysis of composite samples revealed distinct variation in microbial composition (Figs. 5, S7). In the AICB 1382-systems (EC/EZC), several phyla – NB1-j, Cyanobacteria, Gemmatimonadota, Acidobacteriota, Verrucomicrobiota, and Planctomycetota – were primarily observed. These were either underrepresented or absent in the effluent samples. Additional heterotrophic phyla such as Summerlaeota and WPS-2 (Eremiobacterota) were commonly associated. Deinococcota which appeared sporadically in the effluent, Dependentiae phylum, known for its intracellular lifestyle26 and Patescibacteria characterized by minimal genomes27 and epibiotic growth were better represented in AICB 1382 trials.Fig. 5Heatmap (scaled by row) of the first 35 phyla relative abundance in the total biomass collected from the effluent (E), EZ, EC, and EZC (tanks A and B) from E1, E2, and E3 experiments.Full size imageConversely, Proteobacteria (now Pseudomonadota), Bacteroidota, and Actinobacteriota were dominant across all treatments but were most abundant in effluent and EZ (zeolite only) conditions.Approximately 50% of the phyla present in effluents were not detected after culturing. These lost taxa included several anaerobic and extremophilic groups such as Crenarchaeota, Euryarchaeota, Nanoarchaeota, Halobacteriota, Margulisbacteria, Elusimicrobiota, Desulfobacterota, and Fibrobacterota. Additionally, human-associated or potentially pathogenic families belonging to Proteobacteria, Actinobacteria, Campylobacterota, Fusobacteriota, Synergistota, Bacillota (formerly Firmicutes), and Spirochaetota were observed primarily in the effluents, but they were almost absent in the layers dominated by AICB 1382 (Fig. S8). Most genera detected showed abundances below 1% (e.g., Enterobacter spp., Escherichia–Shigella spp., Rickettsia spp., Corynebacterium spp., Mycoplasma spp., and Lachnoclosterium spp.). However, some genera such as Closterium spp., Acinetobacter spp., Pseudomonas spp., Legionella spp., Aeromonas spp., and Mycobacterium spp. accounted for at least 1% of the microbial community when present. The largest values were encountered for Legionella spp. and Clostridium spp. (cca. 4%), and Pseudomonas spp. (cca. 18%). These genera were also identified in the same effluent in a previous study, where they were associated with a high prevalence of antibiotic resistance genes25. The presence of Legionella (7 OTUs), Clostridium (8 OTUs), Pseudomonas (10 OTUs), Acinetobacter (10 OTUs) and Mycobacterium (8 OTUs) genera in the WWs could represent a potential health risk once they enter the receiving rivers, as they are considered important waterborne pathogens25. The presence of the oceanic and hydrothermal vent-associated phylum SAR32428 occurred in the effluent samples.Analysis of the eukaryotic communityComposite samples analysisThe eukaryotic community composition was driven by the same factors, i.e. culture conditions and the effluent microbial load. Unlike the prokaryotic community, the structure of the eukaryotic community was affected more by the effluent than by the treatment applied. PCoA of the composite samples revealed only partial separation according to the applied treatment (Fig. S9). However, clustering analysis (Fig. 6) provided greater resolution, distinguishing the EC/EZC samples from the first and third experiments from those of the second experiment, suggesting a potential effluent-driven grouping. OTU composition showed differences between effluents, with E2E showing the highest diversity (200 OTUs), compared to E1E (102 OTUs) and E3E (25 OTUs) (Fig. S10). At the end of the experiments, the EZC trials retained from the effluent a greater number of taxa than EZ/EC combinations (Fig. S11A−C). Similar to the prokaryotes, each sample harbored a set of distinct eukaryotic taxa (Fig. S11D−F).Fig. 6UPGMA clustering of the biomass samples from the E1, E2, and E3 experiments based on the OTUs abundance, using Kulczynski distance. Each experiment included the effluent (E), EZ, and A and B duplicates of EZC. The E1 and E2 tested the EC condition.Full size imageOverall, alpha diversity indices showed a decline in the species richness during culturing and a similar pattern of the species diversity throughout all samples, except for E3EZC, where it slightly increased, probably due to the larger abundance (Fig. S12A−C). This condition exhibited the highest diversity and abundance indices, along with the lowest Dominance_D, indicating that effluent type played a key role in shaping the community.Three-point samples analysisThe ANOVA assay of the samples collected from the three layers revealed no significant differences between groups (p ≥.001), regardless of experiment, culture condition, or sampling point (Fig. 7A−D). Beta-diversity assay by PCoA and UPGMA (Figs. 8, S13,) further demonstrated that clustering was influenced by the effluent composition and condition tested. In particular, all samples from the EZ trial clustered together; however, the AICB 1382–containing groups separated by effluent type, with E1 and E3 samples distinct from those of the E2 experiment (Fig. 8). The sampling point did not influence the grouping.Fig. 7The species richness, evenness, dominance and diversity indices based on the OTUs abundances from the eukaryotic samples. The samples were grouped by culture conditions EZ, EC, EZC and sampling points (top (1), middle (2) and the bottom layer (3)) (C, D), and by experiment (E1, E2, E3) (A, B). E = effluent samples. The colour of the bars represents group affiliation (shown in the upper right corner of each chart), while the gradient of the bullets indicates variation in the second variable analysed (shown in the lower right corner of each chart).Full size imageFig. 8UPGMA hierarchical clustering of the samples (colour-coded) collected from the three distinct layers: top (1), middle (2), and bottom (3), based on the OTUs abundance using Kulczynski distance. Each experiment included the effluent (E), EZ, and A and B duplicates of EZC. The E1 and E2 included the EC conditi.Full size imageSIMPER analysis showed 93.94% average dissimilarity between the four main clades (Table S7). The top 10 OTUs accounted for 49.44% of overall dissimilarity, primarily including Bacillariophyceae, Xanthophyceae, Cryptomycota, and other heterotrophic taxa.Phylum-level analysisThe analysis of the top 35 most abundant eukaryotic phyla in composite samples and across the three sampling points (Figs. 9, S13) revealed differences among the three effluents (E1E, E2E, and E3E) and culturing conditions (EZ, EC, and EZC). The E1E effluent was dominated by taxa belonging to Stramenopiles – frequently found in urban wastewater29 and Cryptophyceae, known for thriving in diverse environments30 Most of these taxa disappeared during culturing as well as the predators (by myzocystosis) (subphylum Protalveolata)31 and anaerobic phagotrophs (MAST-12 group (Opalomonadea))32 which did not persist post-cultivation. Similarly, in E2E, dominant metazoans like Annelida, Platyhelminthes, Mollusca, and Cnidaria, saprotrophs and parasites from Hyphochytridiomycota33, soil fungi from Basidiomycota and LKM1534, were lost during cultivation. None of these phyla, except for Protalveolata, were identified in the E3E eukaryotic community. This effluent stood out by its abundance of free-living protists from the Centrohelida phylum, found in most aquatic benthic environments where they feed on bacteria and other protists35, and the species-rich Euglenozoa phylum, which contains free-living, parasitic, heterotrophic, and photosynthetic organisms36.Fig. 9Heatmap (scaled by row) of the first 35 eukaryotic phyla sorted according to their relative abundance in the biomass collected from the effluent (E), EZ, EC, and A and B duplicates of EZC. Data Availability. The 16 S rDNA and ribulose-1,5-bisphosphate carboxylase/oxygenase large subunit (rbcL) gene sequences generated during the current study are available in GenBank database with the following IDs PV521982 (https://www.ncbi.nlm.nih.gov/nuccore/PV521982) and PV533918 (https://www.ncbi.nlm.nih.gov/nuccore/PV533918). The 16 S/18S rDNA amplicon datasets generated during the current study are available from the corresponding author on reasonable request. Trichormus variabilis AICB 1382 strain was deposited in the AICB Culture Collection and is available from the corresponding author on reasonable request.Full size imageAn interesting observation was the presence of some taxa at the end of the experiments but their absence at the start in the corresponding effluents (i.e. E1E and E3E), such as Euglenozoa and Centrohelida phyla, in the E1EZ biomass, (Fig. 9), likely the result of sequencing limitations in detecting rare or low-abundance organisms37. Additionally, saprobic, chitin, and keratin-degrading chytrids from Chytridiomycota, which can occasionally act as parasites38, were also identified. Although not present in the E3E community, taxa from the phyla Streptophyta (Viridiplantae), Phragmoplastophyta, and small meiobenthic worm- or cone-shaped animals from Gastrotricha that occur in high abundances in freshwater, marine, and brackish environments39 were identified in high abundance in the EZ sample cultured in this effluent (Figs. 9, S13).In the tested trials (EZ/EC/EZC), green algae (unassigned Chloroplastida) and diatoms (Bacillaryophyta) proliferated in all EZ samples regardless of the effluent type with co-occurring microbial phyla NB1-j, but they were sporadically present in the EC/EZC treatments. In EC/EZC cultures, the dominant phyla were primarily heterotrophs and decomposers. Dominant groups included Labyrinthulomycetes, typically marine saprotrophs or parasites40, free-living heterotrophic protists from phylum Rigifilida41, and fungi like Ascomycota and Cryptomycota42.E1EZC stood out due to its diverse and abundant eukaryotic community, including saprotrophs and parasites soil-fungi from Blastochlamidiomycota43, plant-interacting fungi from Mucoromycota44, molds that feed on bacteria from Fonticula45, and amoeboid taxa like Heterolobosea, biflagellated protists from soil and aquatic habitats from Ancyromonadida46, and free-living amoebae from soil and freshwater Nucleariidae47. These were joined by consumer phyla such as Cercozoa, Ciliophora, and Rotifera, commonly found in urban WW, particularly during warmer seasons29.The clustering analysis (Fig. S14) did not reveal a consistent grouping based on treatment or sampling point. The E2E effluent was more distantly placed compared to E1E and E3E, which intermixed with the cultured samples.DiscussionCommunity composition in engineered aquatic systems was shaped by interacting factors—zeolites, AICB 1382, and effluent type—causing spatial stratification. Due to its buoyant properties, T. variabilis AICB 1382 induced vertical stratification and formed two layers. The upper biofilm layer likely formed due to extracellular polymeric substances (EPS), which vary with conditions and microbial interactions13. Cyanobacterial EPS and oxygen promote bacterial growth, while bacterial metabolism supports algae via CO₂ release and organic matter breakdown48. The intense blue color in the EZC biomass suggests phycobiliprotein (phycocyanin, allophycocyanin) accumulation, sensitive to nitrogen49. Zeolites likely enhanced this by adsorbing and slowly releasing NH₄⁺, reducing volatilization and maintaining nutrient supply50. This localized NH₄⁺ may have supported the cyanobacterial layer at the zeolite surface, as NH₄⁺ is energetically preferred over NO₃⁻51.Effluent type and culturing conditions significantly impacted nutrient removal rate. EC/EZC systems showed the highest PO43− and NO3−/NH4+ removal rate, while EZ performed best with E3E effluent, indicating an effluent-specific effect. Nitrogen recovery matched or exceeded reported rates for Anabaena subcylindrica (19.6–80%) and Nostoc muscorum (20.9–96%)52,53,54. Phosphate removal rate was also comparable or superior to values for A. subcylindrica, N. muscorum (50–81%), Phormidium sp. (62%)55, and Arthrospira sp., which also reduced NH₄⁺ from 100 mg L⁻¹ to < 1 mg L⁻¹56,57. The enhanced performance in cyanobacteria-containing setups likely stemmed from consortia formation between AICB 1382 and native microbiota, known to boost nutrient recovery58. Although zeolite-enhanced systems yielded higher biomass, the observed productivity remains below potential levels reported in literature. For instance, Anabaena sp. reached 720 mg L⁻¹ day⁻¹ in synthetic medium and 400 mg L⁻¹ day⁻¹ in diluted pig slurry, with associated nitrogen removal rates of up to 2471 mg m⁻² day⁻¹59. This suggests that nutrient limitation constrained the biomass accumulation. Future optimisation efforts should consider macronutrient supplementation or the use of nutrient-rich influents to achieve the full potential of the AICB 1382 strain.Effluent composition had the strongest effect on eukaryotic communities, outweighing the influence of AICB 1382 or culturing design. Clustering analysis showed no consistent grouping by treatment or sampling stage, with E2E’s distinct position highlighting effluent chemistry’s role. The difference in the effluent’s chemistry may be due to the seasonal and operational changes in WWTP which introduce variability in the microbial populations60,61. Nutrient stoichiometry, especially deviations from the Redfield N/P ratio of 16 (range: 8.2–45.0), strongly influences microbial diversity62,63. Nevertheless, core bacterial groups—Proteobacteria, Bacteroidota, and Actinobacteriota—were consistently present, reflecting their ecological importance and adaptability in freshwater systems64. Interestingly, SAR324—a bacterial phylum common in oceans, especially near hydrothermal vents—was detected in the effluent samples28, likely originating from the activated sludge microbial community of the WWTP.Zeolites acted as slow-release nutrient carriers and colonization surfaces, promoting both autotrophic and heterotrophic taxa. When combined with AICB 1382, they enhanced diversity, likely supporting rare or slow-growing species65. These effects varied with effluent and culture type, underscoring the need to align interventions with environmental conditions. Zeolites’ aluminosilicate structure favored not only Bacillariophyta but other unassigned Chloroplastida growth in the EZ trial, though T. variabilis competition in EZC conditions likely reduced their abundance. The presence of NB1-j phyla66 further supported diatom viability.The spatial heterogeneity induced by AICB 1382 formed microzones favoring functionally distinct taxa64, including bacteriochlorophyll-a and rhodopsin-bearing groups like Myxococcota, Chloroflexi, and Gemmatimonadota in cyanobacteria-rich layers67,68. Cyanobacteria-associated taxa like Summerlaeota and Eremiobacterota were enriched in cyanobacteria-containing systems69,70, while the presence of resilient Deinococcota71 underscored the selective pressures of engineered environments. N2-fixing cyanobacteria also supported the growth of Verrucomicrobiota, Planctomycetota, and NB1-j, all key players in nutrient cycling and organic matter transformation66,71,72.EC/EZC trials enriched by T. variabilis’ photosynthesis, supported diverse eukaryotic heterotrophs and decomposers commonly associated with primary producers. Elevated biodiversity in samples like E3EZC may reflect higher organic matter from intensified photosynthesis, fostering trophic complexity, or may result from native microbiota or effluent-specific nutrient profiles.Several prokaryotic and eukaryotic phyla declined or disappeared, likely due to oxygenation, competition, or suppression by cyanobacterial metabolites73, suggesting a possible sanitizing effect of the three-layer culturing systems. Lost taxa included extremophiles from salt lakes, intestines, anoxic sediments, and sludge digesters (e.g., Crenarchaeota, Euryarchaeota, Nanoarchaeota, Halobacteriota, Margulisbacteria, Elusimicrobiota, Desulfobacterota)74,75,76,77,78, and gut-associated Fibrobacterota79. Most importantly, detected pathogens form Campylobacterota (formerly Epsilonproteobacteria)80;, Fusobacteriota81; Synergistota, linked to human disease and found in WW, soil, and wells82; Bacillota (Firmicutes)83; and Spirochaetota, also84 were not found at the end of the experiments. Genera like Acinetobacter (Proteobacteria, Moraxellaceae) and Mycobacterium (Actinobacteria) have been positively correlated with the prevalence of carbapenemase-encoding genes which are critical antibiotic resistance determinants25. The occurrence of potential pathogens in effluents underscores WW-related microbial risks, while their elimination in this three-layer system highlights its promise as a future biotechnological approach for WW sanitation and microbial risk mitigation.Conclusion and perspectivesMost harmful bacterial phyla were reduced or eliminated after 14 days of culturing, likely due to oxygen exposure and the allelopathic effects of T. variabilis AICB 1382. This suggests the zeolite–AICB 1382 system poses minimal environmental risk for agricultural use or discharge into surface waters.The AICB 1382 strain dominated EC/EZC treatments, reducing bacterial diversity, whereas zeolites helped maintain higher microbial diversity, particularly in the bottom layer. The three-point sampling revealed distinct microbial stratification, with culturing conditions having the strongest impact on community composition, followed by effluent properties. Zeolites facilitated spatial separation of AICB 1382, contributing to this stratification.Overall, the culture system demonstrated that zeolites and AICB 1382 could modulate the eukaryotic community structure, but the effluent’s chemical background largely dictated the trajectory and clustering of eukaryotic taxa.In conclusion, microbial community dynamics in these systems emerge from the synergistic and antagonistic interplay between effluent characteristics, spatial configuration, and engineered interventions. Zeolites and T. variabilis served as both structural and biological agents capable of shaping ecological outcomes – from diversity reduction to functional specialization. This study demonstrates that combining stratification, nutrient modulation, and bioaugmentation can optimize microbial ecosystems for nutrient recovery, pathogen control, and ecological resilience. Harvested cyanobacterial biomass can be applied in agriculture or safely discharged, while the remaining biomass and zeolites can be reused to inoculate subsequent batches. Applications of the biomass on plants will be addressed in a forthcoming study.”Materials and methodsStrain selection and inoculum preparationThirty xenic cyanobacterial strains from the Algal and Cyanobacterial Culture Collection (AICB), Cluj-Napoca85, were screened for N₂ fixation, growth in effluent, and cell aggregation. Axenic cultures were not used, as the non-sterile effluent and outdoor conditions make contamination unavoidable. Cultures were grown in nitrogen-free BG11 medium86 under natural light (southern exposure) at 19 ± 2 °C. Biomass was sampled for DNA and microscopy, with two transfers into fresh medium to reach exponential growth before inoculation into effluent. Strain AICB 1382 was selected based on its filament aggregation and dark blue color in effluent. Biomass was harvested (4000 rpm, 7 min), weighed, and used in experiments.Light microscopy and taxonomic affiliationMorphological analysis was done using light and fluorescence microscopy with a Nikon TE-2000 Eclipse microscope, and images were captured with a Nikon D90 camera. DNA was extracted using Quick-DNA™ Fecal/Soil Microbe Kits (Zymo Research, Irvine, CA, USA), following the manufacturer’s protocol. PCR and sequencing targeted 16 S rDNA and rbcL genes using primers from Rudi et al.87 and Frank et al.88. The PCR mix included 1.25 U DreamTaq DNA Polymerase (Fermentas, Canada), 1.5 mM MgCl₂, 0.2 mM dNTPs, and 0.4 µM primers in 50 µl total volume. Amplification was done with a Biometra TGradient cycler under standard conditions. Sequencing was performed by Macrogen Europe BV (The Netherlands), and sequences were deposited in GenBank89 under IDs PV521982 and PV533918. Taxonomic identification as Trichormus variabilis was confirmed via BLAST search89 against the GenBank Core Nucleotide database.Experimental design and sampling assayThe 14-day experiment was performed using three separate effluent batches (E1E, E2E, E3E), resulting in three consecutive experimental runs (E1, E2, and E3). Three treatments were tested: effluent with zeolites (EZ), AICB 1382 cultured in effluent with zeolites (EZC), and AICB 1382 cultured in effluent without zeolites (EC). In the E3 run, the EC treatment was omitted due to logistical constraints. Each treatment was applied in a single container, except for EZC, which included two containers labelled A and B. Glass containers (30 × 19 × 20 cm) were kept at 23 ± 2 °C under a 16:8-h light/dark cycle with fluorescent light (25 µmol m⁻² s⁻¹) and placed near a window (northern exposure) to enhance natural illumination. The final irradiance ranged from 40 to 50 µmol m⁻² s⁻¹, varying with weather conditions (sunny versus cloudy days). No stirring was applied to prevent filament breakage55. Unfiltered effluent (1.780 L) originating from the activated sludge process of the municipal wastewater treatment plant in Cluj County, Romania, was collected in September (E1E), August (E2E), and July (E3E). Zeolites (271 g, 3–5 mm, Zeolites Production, Brașov, Romania) were added in EZ and EZC conditions, forming a 0.5 cm layer across 570 cm². AICB 1382 inoculum (500 mg wet biomass) was added to each condition; in EZC, it was mixed with zeolites and layered before pouring the effluent (4 cm liquid hight).Sampling was performed in duplicate at both the beginning and the end of each experiment. At the start, 50 mL samples were collected from the three effluents (E1E, E2E, and E3E). At the end of each experiment, four samples were taken from each culture vessel; for this procedure, 50 mL were collected by pipetting separately from the upper and middle layers, without mixing them. For the bottom (zeolite) layer, an equivalent of 50 ml was estimated based on the total effluent volume (1.780 L) and zeolite weight (271 g), resulting in 7.5 g of zeolites. These were rinsed with 10 ml of a MgSO₄·7 H₂O (10 mM) and Tween 80 (2000:1 v/v) solution to detach biofilm. After separate sampling, the contents were mixed thoroughly, and a final 50 ml composite sample was taken. All samples were filtered through sterile 0.22 μm cellulose nitrate membranes (Sartorius); filtrates were reserved for nutrient analysis. Filters were weighed before and after filtration to determine biomass, then stored at − 20 °C.Biomass and nutrient analysisWet biomass yield was calculated by summing biomass from all sampling points and subtracting the 500 mg inoculum. Nutrient levels were measured using HANNA Instruments kits with a HAN I83399 Multiparameter Photometer (HANNA Instruments, Germany): PO₄³⁻ (kit HI 93713-01), and NO₃⁻/NH₄⁺ (kit HI 93767 A-50), reported in mg L⁻¹.16 S/18S rRNA gene metagenomic sequencingNucleic acids were extracted from membranes using the kit described in Sect. 5.2, with duplicates pooled. PCR, quality control, amplicon library preparation, and sequencing were performed by Novogene CO using Illumina PE250 (30 K tags/sample). The bacterial 16 S rRNA V3–V4 region was amplified with primers 341 F/806R90, and the eukaryotic 18 S rRNA V4 region with primers 528 F/706R91. Reads were demultiplexed, barcodes/primers removed, and merged with FLASH92. Quality filtering followed QIIME 293; chimeras were removed using UCHIME. OTUs were clustered at 97% similarity via UPARSE94 Taxonomic classification used Mothur (archaea/bacteria) and RDP (eukaryotes) with the SILVA database v138.195.Diversity, statistical analysis, and visual representation of taxaTo reduce experimental error and ensure comparability, OTU abundances were normalized to the sample with the fewest sequences. Alpha diversity was assessed using Chao-1, Shannon_H, Evenness e^H/S, and Dominance_D indices. Beta diversity was analyzed via SIMPER, PCoA, and UPGMA clustering using Bray–Curtis and Kulczynski distances. Analyses were done in PAST 4.1396, and one-way ANOVA was performed in JASP v.0.19.3 (2025). OTU visualization via Venn and flower plots used EVenn97, and heatmaps were created in TBtools-II98.

    Data availability

    The 16 S rDNA and ribulose-1,5-bisphosphate carboxylase/oxygenase large subunit (rbcL) gene sequences generated during the current study are available in GenBank database with the following IDs PV521982 ([https://www.ncbi.nlm.nih.gov/nuccore/PV521982](https:/www.ncbi.nlm.nih.gov/nuccore/PV521982)) and PV533918 ([https://www.ncbi.nlm.nih.gov/nuccore/PV533918](https:/www.ncbi.nlm.nih.gov/nuccore/PV533918)).The 16 S/18S rDNA amplicon datasets generated during the current study are available from the corresponding author on reasonable request.*Trichormus variabilis* AICB 1382 strain was deposited in the AICB Culture Collection and is available from the corresponding author on reasonable request.
    ReferencesXiao, J. et al. Integrating desertification control and wastewater treatment: novel insights from the induction of artificial biocrusts using municipal wastewater-cultivated Cyanobacterium. Sci. Total Environ. 955, 177049. https://doi.org/10.1016/j.scitotenv.2024.177049 (2024).Article 
    CAS 
    PubMed 

    Google Scholar 
    Haraguchi, Y. & Shimizu, T. Crop cultivation without nitrogen fertiliser using nitrogen-fixing cyanobacterial extracts for low environmental impact. Sci. Rep. 15, 18365. https://doi.org/10.1038/s41598-025-01741-5 (2025).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kumar, R., Singh, S. & Sharma, R. C. Application of WQI for assessment of water quality of high altitude lake Dodi Tal, Garhwal Himalaya, India. Sustain. Water Resour. Manage. 5, 1033–1042. https://doi.org/10.1007/s40899-018-0281-1 (2019).Article 

    Google Scholar 
    Bisht, N., Chauhan, P.S. Excessive and disproportionate use of chemicals cause soil contamination and nutritional stress in Soil Contamination-Threats and Sustainable Solutions. (eds. Larramendy, M.L., Soloneski, S.) 1–10 (IntechOpen, 2020).Sánchez-Quintero, Á., Fernandes, S. C. M. & Beigbeder, J. B. Overview of microalgae and cyanobacteria-based biostimulants produced from wastewater and CO2 streams towards sustainable agriculture: A review. Microbiol. Res. 277, 127505. https://doi.org/10.1016/j.micres.2023.127505 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Moore, A. W. Azolla: biology and agronomic significance. Bot. Rev. 35, 17–34. https://doi.org/10.1007/BF02859886 (1969).Article 
    CAS 

    Google Scholar 
    Yadav, P. et al. Mechanisms of stress tolerance in cyanobacteria under extreme conditions. Stresses 2, 531–549. https://doi.org/10.3390/stresses2040036 (2022).Article 

    Google Scholar 
    Ahmad, I. Z. The usage of cyanobacteria in wastewater treatment: prospects and limitations. LAM 75, 718–730. https://doi.org/10.1111/lam.13587 (2022).Article 
    CAS 
    PubMed 

    Google Scholar 
    Kar, J. et al. Revisiting the role of cyanobacteria-derived metabolites as antimicrobial agent: A 21st century perspective. Front. Microbiol. 13, 1034471. https://doi.org/10.3389/fmicb.2022.1034471 (2022).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Ferreira, A. et al. Biostimulant and biopesticide potential of microalgae growing in piggery wastewater. Environ. Adv. 4, 100062. https://doi.org/10.1016/j. envadv.2021.100062 (2021).Article 
    CAS 

    Google Scholar 
    Bellver, M., Díez-Montero, R., Escol`a Casas, M., Matamoros, V. & Ferrer, I. Phycobiliprotein recovery coupled to the tertiary treatment of wastewater in semi-continuous photobioreactors. Tracking contaminants of emerging concern. Bioresour Technol. 384, 129287. https://doi.org/10.1016/j.biortech.2023.129287 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Bhasin, A., Hussain, A. S. & Simsek, H. Biodegradability and bioavailability of dissolved substances in aquaculture effluent: performance of Indigenous bacteria, cyanobacteria, and green microalgae. Environ. Pollut. 345, 123468. https://doi.org/10.1016/j.envpol.2024.123468 (2024).Article 
    CAS 
    PubMed 

    Google Scholar 
    Nishanth, S. et al., Cyanobacterial extracellular polymeric substances (EPS): Biosynthesis and their potential applications in Microbial and Natural Macromolecules. (eds. Das, S., Dash, H.R.) 349–369 (Academic Press, 2021).Ramanan, R., Kim, B. H., Cho, D. H., Oh, H. M. & Kim, H. S. Algae–bacteria interactions: evolution, ecology and emerging applications. Biotechnol. Adv. 34, 14–29. https://doi.org/10.1016/j.biotechadv.2015.12.003 (2015).Article 
    CAS 
    PubMed 

    Google Scholar 
    Fang, Y., Lin, G., Liu, Y. & Zhang, J. Advanced treatment of antibiotic-polluted wastewater by a consortium composed of bacteria and mixed cyanobacteria. Environ. Pollut. 344, 123293. https://doi.org/10.1016/j.envpol.2024.123293 (2024).Article 
    CAS 
    PubMed 

    Google Scholar 
    Abed, R. M. M. Interaction between cyanobacteria and aerobic heterotrophic bacteria in the degradation of hydrocarbons. Int. Biodeterior. Biodegrad. 64, 58–64. https://doi.org/10.1016/j.ibiod.2009.10.008 (2010).Article 
    CAS 

    Google Scholar 
    Munoz, R. & Guieysse, B. Algal–bacterial processes for the treatment of hazardous contaminants: a review. Water Res. 40, 2799–2815. https://doi.org/10.1016/j.watres.2006.06.011 (2008).Article 
    ADS 
    CAS 

    Google Scholar 
    Tchobanoglous, G., Burton, F.L. & Stensel, H.D., Wastewater Engineering: Treatment and Reuse. (eds. Tchobanoglous, G., Burton, F.L., Stensel, H.D.). 4th edition (McGraw-Hill, 2002).Talukder, A., Mahmud, S., Lira, S. & Aziz, M. Phycoremediation of textile industry effluent by cyanobacteria (Nostoc muscorum and Anabaena variabilis). Biores Commun. 1, 124–127 (2015).
    Google Scholar 
    Barrer, R. M. Zeolites and Clay Minerals as Sorbents and Molecular Sieves. (Academic Press, 1978).Markou, G., Vandamme, D. & Muylaert, K. Using natural zeolite for ammonia sorption from wastewater and as nitrogen releaser for the cultivation of arthrospira platensis. Bioresour Technol. 155, 373–378. https://doi.org/10.1016/j.biortech.2013.12.122 (2014).Article 
    CAS 
    PubMed 

    Google Scholar 
    Blanchard, G., Maunaye, M. & Martin, G. Removal of heavy metals from waters by means of natural zeolites. Water Res. 18, 1501–1507. https://doi.org/10.1021/es048482s (1984).Article 
    ADS 
    CAS 

    Google Scholar 
    Kubota, M. et al. Selective adsorption of bacterial cells onto zeolites, colloids surf. B Biointerfaces. 64, 88–97. https://doi.org/10.1016/j.colsurfb.2008.01.012 (2008).Article 
    CAS 
    PubMed 

    Google Scholar 
    Emami Moghaddam, S. A., Harun, R., Mokhtar, M. N. & Zakaria, R. Potential of zeolite and algae in biomass immobilization. Biomed. Res. Int. https://doi.org/10.1155/2018/6563196 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Teban-Man, A. et al. Municipal wastewaters carry important carbapenemase genes independent of hospital input and can mirror clinical resistance patterns. Microbiol. Spectr. https://doi.org/10.1128/spectrum.02711-21 (2022).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Weisse, L., Héchard, Y., Moumen, B. & Delafont, V. Here, there and everywhere: ecology and biology of the dependentiae phylum. Environ. Microbiol. 25, 597–605. https://doi.org/10.1111/1462-2920.16307 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Castelle, C. J. et al. Biosynthetic capacity, metabolic variety and unusual biology in the CPR and DPANN radiations. Nat. Rev. Microbiol. 16, 629–645. https://doi.org/10.1038/s41579-018-0076-2 (2018).Article 
    CAS 
    PubMed 

    Google Scholar 
    Malfertheiner, L., Martínez-Pérez, C., Zhao, Z., Herndl, G. J. & Baltar, F. Phylogeny and metabolic potential of the candidate phylum SAR324. Biology (Basel). 11, 599. https://doi.org/10.3390/biology11040599 (2022).Article 
    CAS 
    PubMed 

    Google Scholar 
    Maritz, J. M., Eyck, T., Alter, T. A., Carlton, J. M. & E.S. & Patterns of protist diversity associated with Raw sewage in new York City. ISME J. 13, 2750–2763. https://doi.org/10.1038/s41396-019-0467-z (2019).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Qu, F., Wang, Y., Yu, D. & Chen, N. High-frequency monitoring reveals phytoplankton succession patterns and the role of cryptophyte in a subtropical river reservoir. Algal Res. 82, 103680. https://doi.org/10.1016/j.algal.2024.103680 (2024).Article 

    Google Scholar 
    Cavalier-Smith, T. & Chao, E. E. Protalveolate phylogeny and systematics and the origins of sporozoa and dinoflagellates (phylum myzozoa nom. Nov). Eur. J. Protistol. 40, 185–212. https://doi.org/10.1016/j.ejop.2004.01.002 (2004).Article 

    Google Scholar 
    Cavalier-Smith, T. & Scoble, J. M. Phylogeny of heterokonta: Incisomonas marina, a uniciliate gliding opalozoan related to solenicola (Nanomonadea), and evidence that actinophryida evolved from raphidophytes. Eur. J. Protistol. 49, 328–353. https://doi.org/10.1016/j.ejop.2012.09.002 (2012).Article 
    PubMed 

    Google Scholar 
    Beakes, G. W. & Thines, M. In Hyphochytriomycota and Oomycota in Handbook of the Protists (eds Archibald, J. M. et al.) 435–505 (Springer, 2017).Chapter 

    Google Scholar 
    Zhang, L. et al. Immediate and legacy effects of snow exclusion on soil fungal diversity and community composition. Ecosyst. 8, 22. https://doi.org/10.1186/s40663-021-00299-8 (2021).Article 

    Google Scholar 
    Gast, R. J. Centrohelida and other Heliozoan-like protists Handbook of the Protists. (eds. Archibald, J.M., Simpson, G.B.A. & Slamovits, C.H.) 955–971 (Springer, 2017). (2017).Kostygov, A. Y. et al. Euglenozoa: taxonomy, diversity and ecology, symbioses and viruses. Open. Biol. 11, 200407. https://doi.org/10.1098/rsob.200407 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Jia, Y. et al. Sequencing introduced false positive rare taxa lead to biased microbial community diversity, assembly, and interaction interpretation in amplicon studies. Environ. Microbiome. 17, 43. https://doi.org/10.1186/s40793-022-00436-y (2022).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Sparrow, F. K. Aquatic Phycomyete. (2nd ed.) (ed. Arbor, A.) 42–119 (The University of Michigan Press, 1960).Wey-Fabrizius, A. R., Podsiadlowski, L., Herlyn, H. & Hankeln, T. Platyzoan mitochondrial genomes. Mol. Phylogenet Evol. 69, 365–375. https://doi.org/10.1016/j.ympev.2012.12.015 (2013).Article 
    PubMed 

    Google Scholar 
    Tsui, C. K. et al. Labyrinthulomycetes phylogeny and its implications for the evolutionary loss of chloroplasts and gain of ectoplasmic gliding. Mol. Phylogenet Evol. 50, 129–140. https://doi.org/10.1016/j.ympev.2008.09.02 (2009).Article 
    CAS 
    PubMed 

    Google Scholar 
    Yabuki, A., Ishida, K. & Cavalier-Smith, T. Rigifila Ramosa n. gen., n. sp., a filose apusozoan with a distinctive pellicle, is related to micronuclearia. Protist 164, 75–88. https://doi.org/10.1016/j.protis.2012.04.005 (2013).Article 
    PubMed 

    Google Scholar 
    Naranjo-Ortiz, M. A. & Gabaldón, T. Fungal evolution: diversity, taxonomy, and phylogeny of the fungi. Biol. Rev. Camb. Philos. Soc. 94, 2101–2137. https://doi.org/10.1111/brv.12550 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    McConnaughey, M. In Physical chemical properties of Fungi in Reference Module in Biomedical Sciences (ed. Caplan, M. J.) 1–4 (Elsevier, 2014).
    Google Scholar 
    Bonfante, P. & Venice, F. Mucoromycota: going to the roots of plant-interacting fungi. Fungal Biol. Rev. 34, 100–113. https://doi.org/10.1016/j.fbr.2019.12.003 (2020).Article 

    Google Scholar 
    Toret, C., Picco, A., Boiero-Sanders, M., Michelot, A. & Kaksonen, M. The cellular slime mold fonticula Alba forms a dynamic, multicellular collective while feeding on bacteria. Curr. Biol. 32, 1961–1973. https://doi.org/10.1016/j.cub.2022.03.018 (2022).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Cavalier-Smith, T. Early evolution of eukaryote feeding modes, cell structural diversity, and classification of the protozoan phyla Loukozoa, Sulcozoa, and choanozoa. Eur. J. Protistol. 49, 115–178. https://doi.org/10.1016/j.ejop.2012.06.001 (2013).Article 
    PubMed 

    Google Scholar 
    Zettler, N. T., O’Kelly, C. & Sogin, M. The Nucleariid amoebae: more protists at the animal-fungal boundary. J. Eukaryot. Microbiol. 48, 293–297. https://doi.org/10.1111/j.1550-7408.2001.tb00317.x (2001).Article 

    Google Scholar 
    Cuellar-Bermudez, S. P. et al. Nutrients utilization and contaminants removal. A review of two approaches of algae and cyanobacteria in wastewater. Algal Res. 24, 438–449. https://doi.org/10.1016/j.algal.2016.08.018 (2017).Article 

    Google Scholar 
    Simeunović, J., Bešlin, K., Svirčev, Z., Kovač, B. D. & Babić, O. Impact of nitrogen and drought on phycobiliprotein content in terrestrial cyanobacterial strains. J. Appl. Phycol. 25, 597–607. https://doi.org/10.1007/s10811-012-9894-1 (2012).Article 
    CAS 

    Google Scholar 
    Fernandes, F. A. et al. Using the von Bertalanffy model to describe ammonia volatilization from conventional and stabilized nitrogen fertilizers in coffee crop system. Commun. Soil. Sci. Plant. Anal. 55, 1837–1848. https://doi.org/10.1080/00103624.2024.2330628 (2024).Article 
    CAS 

    Google Scholar 
    Collier, L. J., Lovindeer, R., Xi, Y., Radway, C. J. & Armstrong, A. R. Differences in growth and physiology of marine Synechococcus (Cyanobacteria) on nitrate versus ammonium are not determined solely by nitrogen source redox state. J. Phycol. 48, 106–116. https://doi.org/10.1111/j.1529-8817.2011.01100.x (2012).Article 
    CAS 
    PubMed 

    Google Scholar 
    El-Sheekh, M. M., El-Shouny, W. A., Osman, M. E. & El-Gammal, E. Treatment of sewage and industrial wastewater effluents by the cyanobacteria Nostoc muscorum and Anabaena subcylinderica. J. Water Chem. Technol. 36, 190–197. https://doi.org/10.3103/S1063455X14040079 (2014).Article 

    Google Scholar 
    Mallick, N. & Rai, L. C. Removal of inorganic ions from wastewater by immobilized microalgae. World J. Microbiol. Biotechnol. 10, 439–443. https://doi.org/10.1007/BF00144469 (1994).Article 
    CAS 
    PubMed 

    Google Scholar 
    Deb, D., Mallick, N. & Bhadoria, P. B. S. A waste-to-wealth initiative exploiting the potential of Anabaena variabilis for designing an integrated biorefinery. Sci. Rep. 12, 9478. https://doi.org/10.1038/s41598-022-13244-8 (2022).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Pouliot, Y., Buelna, G., Racine, C. & de la Noüe, J. Culture of cyanobacteria for tertiary wastewater treatment and biomass production. Biol. Wastes. 29, 81–91. https://doi.org/10.1016/0269-7483(89)90089-X (1989).Article 
    CAS 

    Google Scholar 
    Lincoln, E. P., Wilkie, A. C. & French, B. T. Cyanobacterial process for renovating dairy wastewater. Biomass Bioenergy. 10, 63–68. https://doi.org/10.1016/0961-9534(95)00055-0 (1996).Article 
    CAS 

    Google Scholar 
    Markou, G., Chatzipavlidis, I. & Georgakakis, D. Cultivation of arthrospira (Spirulina) platensis in olive-oil mill wastewater treated with sodium hypochlorite. Bioresour Technol. 112, 234–241. https://doi.org/10.1016/j.biortech.2012.02.098 (2012).Article 
    CAS 
    PubMed 

    Google Scholar 
    Salbitani, G. & Carfagna, S. Ammonium utilization in microalgae: A sustainable method for wastewater treatment. Sustainability 13, 956. https://doi.org/10.3390/su13020956 (2021).Article 
    ADS 
    CAS 

    Google Scholar 
    Morillas-España, A. et al. Potential of the cyanobacteria Anabaena sp. and Dolichospermum sp. for being produced using wastewater or pig slurry: validation using pilot-scale raceway reactors. Algal Res. 60, 102517. https://doi.org/10.1016/j.algal.2021.102517 (2001).Article 

    Google Scholar 
    Alsulaili, A., Al-Buloushi, B. Y. & Hamoda, M. F. Seasonal variation pattern of physicochemical and microbial parameters in a wastewater treatment plant. Desalin. Water Treat. 208, 244–260. https://doi.org/10.5004/dwt.2020.26461 (2020).Article 
    CAS 

    Google Scholar 
    Johnston, J. & Behrens, S. Seasonal dynamics of the activated sludge microbiome in sequencing batch reactors, assessed using 16S rRNA transcript amplicon sequencing. Appl. Environ. Microbiol. https://doi.org/10.1128/AEM.00597-20 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Gao, L. et al. Effects of different nitrogen/phosphorus ratios on the growth and metabolism of microalgae scenedesmus obliquus cultured in the mixed wastewater from primary settling tank and sludge thickener. Process. Saf. Environ. Prot. 170, 824–833. https://doi.org/10.1016/j.psep.2022.12.059 (2023).Article 
    CAS 

    Google Scholar 
    Qian, W. et al. Effect of N/P ratio on attached microalgae growth and the differentiated metabolism along the depth of biofilm. Environ. Res. 240 240, 117428. https://doi.org/10.1016/j.envres.2023.117428 (2024).Article 
    CAS 

    Google Scholar 
    Uthpala Pushpakumara, B. L. D., Tandon, K., Willis, A. & Verbruggen, H. Unravelling microalgal-bacterial interactions in aquatic ecosystems through 16S rRNA gene-based co-occurrence networks. Sci. Rep. 13, 2743. https://doi.org/10.1038/s41598-023-27816-9 (2023).Article 
    ADS 
    CAS 

    Google Scholar 
    Vallina, S. M., Cermeno, P., Dutkiewicz, S., Loreau, M. & Montoya, J. M. Phytoplankton functional diversity increases ecosystem productivity and stability. Ecol. Model. 361, 184–196. https://doi.org/10.1016/j.ecolmodel.2017.06.020 (2017).Article 

    Google Scholar 
    Berry, D. & Widder, S. Deciphering microbial interactions and detecting keystone species with co-occurrence networks. Front. Microbiol. 5, 1–13. https://doi.org/10.3389/fmicb.2014.00219 (2014).Article 

    Google Scholar 
    Vigneron, A. et al. Multiple strategies for light-harvesting, photoprotection, and carbon flow in high latitude microbial Mats. Front. Microbiol. 9, 2881. https://doi.org/10.3389/fmicb.2018.02881 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Mujakić, I. et al. Multi-environment ecogenomics analysis of the cosmopolitan phylum Gemmatimonadota. Microbiol. Spectr. https://doi.org/10.1128/spectrum.01112-23 (2023).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Fang, Y. et al. Casting light on the adaptation mechanisms and evolutionary history of the widespread Sumerlaeota. mBio 12, e00350-2. https://doi.org/10.1128/mbio.00350-21 (2021).Article 
    CAS 

    Google Scholar 
    Sheremet, A. et al. Ecological and genomic analyses of candidate phylum WPS-2 bacteria in an unvegetated soil. Environ. Microbiol. 22, 3143–3157. https://doi.org/10.1111/1462-2920.15054 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Martinez-Garcia, M. et al. Capturing single cell genomes of active polysaccharide degraders: an unexpected contribution of verrucomicrobia. PLoS ONE. 7, e35314. https://doi.org/10.1371/journal.pone.0035314 (2012).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bellini, E. et al. Trichormus variabilis (Cyanobacteria) biomass: from the nutraceutical products to novel EPS-cell/protein carrier systems. Mar. Drugs. 16, 298. https://doi.org/10.3390/md16090298 (2018).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Yücer, D. T., Beyatlı, Y. & Pabuçcu, K. The antiproliferative and antimicrobial effects of cultivated Anabaena circinalis rabenhorts ex Bornet and Flahault and Nostoc entophytum Bornet and Flahault. Trop. J. Pharm. Res. 17, 1571–1577. https://doi.org/10.4314/tjpr.v17i8.15 (2018).Article 
    CAS 

    Google Scholar 
    Pesaro, M. & Widmer, F. Identification of novel crenarchaeota and Euryarchaeota clusters associated with different depth layers of a forest soil. FEMS Microbiol. Ecol. 42, 89–98. https://doi.org/10.1111/j.1574-6941.2002.tb00998.x (2002).Article 
    CAS 
    PubMed 

    Google Scholar 
    Stingl, U., Radek, R., Yang, H. & Brune, A. Endomicrobia: cytoplasmic symbionts of termite gut protozoa form a separate phylum of prokaryotes. Appl. Environ. Microbiol. 71, 1473–1479. https://doi.org/10.1128/AEM.71.3.1473-1479.2005 (2005).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Matheus Carnevali, P. B. et al. Hydrogen-based metabolism as an ancestral trait in lineages sibling to the cyanobacteria. Nat. Commun. 10, 463. https://doi.org/10.1038/s41467-018-08246-y (2019).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Murphy, C. L. et al. Genomic characterization of three novel Desulfobacterota classes expand the metabolic and phylogenetic diversity of the phylum. Environ. Microbiol. 23, 4326–4343. https://doi.org/10.1111/1462-2920.15614 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Wittmers, F. et al. Non-photosynthetic lineages sibling to Cyanobacteria associate with eukaryotes in the open ocean. Curr. Biol. 34, R1133–R1134. https://doi.org/10.1100/tsw.2001.16 (2024).Article 
    CAS 
    PubMed 

    Google Scholar 
    Neumann, A. P., McCormick, C. A. & Suen, G. Fibrobacter communities in the Gastrointestinal tracts of diverse hindgut-fermenting herbivores are distinct from those of the rumen. Environ. Microbiol. 19, 3768–3783. https://doi.org/10.1111/1462-2920.13878 (2017).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wösten, M. M. Regulation of respiratory pathways in campylobacterota: A review. Front. Microbiol. 10, 424488. https://doi.org/10.3389/fmicb.2019.01719 (2019). (2019).Article 

    Google Scholar 
    Bullman, S. et al. Analysis of Fusobacterium persistence and antibiotic response in colorectal cancer. Science 358, 1443–1448 (2017). https://doi.org/aal5240Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Jumas-Bilak, E. et al. Jonquetella anthropi gen. nov., sp. nov., the first member of the candidate phylum ‘Synergistetes’ isolated from man. Int. J. Syst. Evol. Microbiol. 57, 2743–2748. https://doi.org/10.1099/ijs.0.65213-0 (2007).Article 
    CAS 
    PubMed 

    Google Scholar 
    Wolf, M., Müller, T., Dandekar, T. & Pollack, J. D. Phylogeny of firmicutes with special reference to Mycoplasma (Mollicutes) as inferred from phosphoglycerate kinase amino acid sequence data. Int. J. Syst. Evol. Microbiol. (Comparative Study). 54, 871–875. https://doi.org/10.1099/ijs.0.02868-0 (2004).Article 
    CAS 

    Google Scholar 
    Margulis, L., Ashen, J. B., Solé, M. & Guerrero, R. Composite, large spirochetes from microbial mats: spirochete structure reviews. PNAS 90, 6966–6970. https://doi.org/10.1073/pnas.90.15.6966 (1993).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Dragoș, N., Péterfi, L.Ș., Momeu, L. & Popescu, C. An introduction to the Algae and the Culture Collection of Algae at the Institute of Biological Research Cluj-Napoca. 1st edition (Cluj University Press, 1997).Stanier, R. Y., Kunisawa, R., Mandel, M. & Cohen-Bazire, G. Purification and properties of unicellular blue-green algae (order Chroococcales). Bacteriol. Rev. 35, 171–205. https://doi.org/10.1128/br.35.2.171-205.1971 (1971).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Rudi, K., Skulberg, O. M. & Jakobsen, K. S. Evolution of cyanobacteria by exchange of genetic material among phyletically related strains. J. Bacteriol. 180, 3453–3461. https://doi.org/10.1128/jb.180.13.3453-3461.1998 (1998).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Frank, J. A. et al. Critical evaluation of two primers commonly used for amplification of bacterial 16S rRNA genes. Appl. Environ. Microbiol. 74, 2461–2470. https://doi.org/10.1128/AEM.02272-07 (2008).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Benson, D. A. et al. GenBank NAR 41, D36–D42 https://doi.org/10.1093/nar/gks1195. (2013).Article 
    CAS 
    PubMed 

    Google Scholar 
    Stibal, M. et al. Different bulk and active bacterial communities in cryoconite from the margin and interior of the Greenland ice sheet. Environ. Microbiol. Rep. 7, 293–300. https://doi.org/10.1111/1758-2229.12246 (2015).Article 
    CAS 
    PubMed 

    Google Scholar 
    Cheung, M. K., Au, C. H., Chu, K. H., Kwan, H. S. & Wong, C. K. Composition and genetic diversity of picoeukaryotes in subtropical coastal waters as revealed by 454 pyrosequencing. ISME J. 4, 1053–1059. https://doi.org/10.1038/ismej.2010.26 (2010).Article 
    ADS 
    PubMed 

    Google Scholar 
    Magoč, T., Salzberg, S. L. & FLASH Fast length adjustment of short reads to improve genome assemblies. Bioinformatics 27, 2957–2963. https://doi.org/10.1093/bioinformatics/btr507 (2011).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bolyen, E. et al. Reproducible, interactive, scalable and extensible Microbiome data science using QIIME 2. Nat. Biotechnol. 37, 852–857. https://doi.org/10.1038/s41587-019-0209-9 (2019).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Edgar, R. C. UNOISE2: improved error-correction for illumina 16S and ITS amplicon sequencing. Preprint at. https://doi.org/10.1101/081257 (2016).Article 

    Google Scholar 
    Quast, C. et al. The SILVA ribosomal RNA gene database project: improved data processing and web-based tools. NAR 41, D590–D596. https://doi.org/10.1093/nar/gks1219 (2013).Article 
    CAS 
    PubMed 

    Google Scholar 
    Hammer, Ø., Harper, D. A. T., Ryan, P. D. & PAST Paleontological statistics software package for education and data analysis. Palaeontol. Electronica. 4, 1–9 (2001). http://palaeo-electronica.org/2001_1/past/issue1_01.htm
    Google Scholar 
    Mei, Y., Tong, C., Yong-Xin, L. & Luqi, H. Visualizing set relationships: EVenn’s comprehensive approach to Venn diagrams. Imeta https://doi.org/10.1002/imt2.184 (2024).Article 

    Google Scholar 
    Chen, C. et al. TBtools-II: A one for all, all for one bioinformatics platform for biological big-data mining. Mol. Plant. 16, 1733–1742. https://doi.org/10.1016/j.molp.2023.09.010 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Download referencesAcknowledgementsWe would like to acknowledge Compania de Apă Someș S.A. for their support in providing the effluent from the Municipal Wastewater Treatment Plant.FundingThis work was supported by the Romanian Ministry of Research, Innovation and Digitization through Nucleu Program under 2022–2027 National Research, Development and Innovation Plan [PN23020401, contract no. 7 N/03.01.2023]; Romanian Ministry of Research, Innovation and Digitization [PN-III-P2-2.1-PED-2021, contract 653/2022]; National Recovery and Resilience Plan (PNRR) [760102/23.05.2023].Author informationAuthors and AffiliationsInstitute of Biological Research Cluj, National Institute of Research and Development for Biological Sciences, 48 Republicii Street, 400015, Cluj-Napoca, RomaniaAdriana Hegedűs, Răzvan Vințan, Maria Nicoară & Bogdan DrugăFaculty of Biology and Geology, “Babeş-Bolyai“ University, 5-7 Clinicilor St., Cluj-Napoca, 400006, RomaniaRăzvan VințanDoctoral School of Integrative Biology, Faculty of Biology and Geology, “Babeş-Bolyai“ University, 44 Republicii Street, Cluj-Napoca, 400015, RomaniaMaria NicoarăAuthorsAdriana HegedűsView author publicationsSearch author on:PubMed Google ScholarRăzvan VințanView author publicationsSearch author on:PubMed Google ScholarMaria NicoarăView author publicationsSearch author on:PubMed Google ScholarBogdan DrugăView author publicationsSearch author on:PubMed Google ScholarContributionsAdriana Hegedűs: Writing—original draft, Methodology, Data curation, Formal analysis. Răzvan Vințan: Methodology, Investigation, Data curation. Maria Nicoară: Methodology, Investigations.Bogdan Drugă: Conceptualization, Funding acquisition, Writing—review & editing, Supervision; Validation.Corresponding authorCorrespondence to
    Bogdan Drugă.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 articleHegedűs, A., Vințan, R., Nicoară, M. et al. Synergistic role of Trichormus variabilis and zeolites in three-layer culturing system for modulating the wastewater effluent community.
    Sci Rep 15, 44176 (2025). https://doi.org/10.1038/s41598-025-27997-5Download citationReceived: 16 September 2025Accepted: 06 November 2025Published: 19 December 2025Version of record: 19 December 2025DOI: https://doi.org/10.1038/s41598-025-27997-5Share 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
    Keywords
    Trichormus variabilis
    ZeolitesNutrientsBiomassProkaryotic communityEukaryotic community More

  • in

    Characterization of the chloroplast genome of a relict tree, Pterocarya fraxinifolia (Juglandaceae), and its comparative analysis

    AbstractThe rare, vulnerable relict species Pterocarya fraxinifolia is among the last surviving tree species growing in small, scattered populations in the southern Caucasus region; P. fraxinifolia grows up to 1000 m in plain forests and is threatened by habitat loss and environmental changes. Here, we sequenced and annotated the chloroplast genome of P. fraxinifolia from Hyrcanian forests and compared it to the chloroplast genomes of five other Pterocarya species. The evolutionary relationships of P. fraxinifolia were subsequently evaluated using the chloroplast genomes and individual chloroplast loci. The chloroplast genome of P. fraxinifolia was 160,086 bp in length, comprising 128 genes and a typical quadripartite structure. A comparative analysis of the six Pterocarya species revealed limited nucleotide diversity and structural variations in genes. The bulk of the 68 loci identified by SSR analysis comprised A/T repeats. Codon bias analysis revealed strong purifying selection, with the ndhF gene showing the highest Ka/Ks ratio. Our phylogenetic analysis revealed Pterocarya as a sister to the genus Juglans and a distinct subclade within Pterocarya.

    IntroductionRelict species have always excited evolutionary biologists and biogeographers who consider these species ‘living fossils’ or relics of prehistoric periods1,2. These species have great value as research models for the geographical distribution of intercontinental rifts and as species that ensure biodiversity and ecosystem balance. Relict species also provide relevant information about the adaptation of species to specific environmental changes, as well as the impact of climate change on the animal and plant kingdoms3.Hyrcanian forests are hotspots for biodiversity and are home to numerous relict species4, including 280 endemic and subendemic species5,6,7. The genus Pterocarya Kunth (Juglandaceae), commonly referred to as wingnuts, has a disjunct distribution in East Asia and the Caucasus region with its most recent common ancestor present 40 Ma8. Pterocarya comprises six species, which are classified into two sections, Pterocarya (P. fraxinifolia, P. hupehensis, P. stenoptera, and P. tonkinensis) and Platyptera (P. macroptera and P. rhoifolia), on the basis of the presence or absence of scales on the terminal buds9. P. fraxinifolia is the only species in western Asia10. The remaining species of Pterocarya occur in eastern Asia, such as China and Japan11,12,13,14,15. Recently, a series of studies have focused on the phylogeny, biogeography, population genetics, and landscape genetics of species in this genus14,15. However, resources regarding the chloroplast genome in this genus are insufficient, and more research is still needed. Pterocarya fraxinifolia is a deciduous tree that can reach 20–25 m in height and 1.8 m in trunk diameter and is wind-pollinated to produce wing-nut fruits12. This species is among the last surviving trees growing in small scattered populations in the southern Caucasus region, which includes northern Iran, Georgia, Armenia, Azerbaijan, and the Anatolian region in Turkey12,15. However, less than two decades ago, small populations were first recorded in western Iran in the provinces of Lorestan and Ilam in the Zagros Mountains16.The chloroplast genome is widely used in phylogenetic studies because of its relatively conserved structure17,18 and uniparental inheritance19,20,. Chloroplast genomes can provide important information about the adaptation of species to different environmental conditions21,22,23. Despite the slow evolutionary rates of chloroplast genomes, coding and noncoding regions are useful for the identification of closely related species24,25,26 and for detecting genome-scale evolutionary patterns. Comparisons of the structure and sequence of these regions across different species within a genus can reveal important evolutionary phenomena such as gene transfer, deletion, or duplication. Recently, with the continuous application of high-throughput sequencing techniques, chloroplast DNA sequences have become readily available13,27,28. However, there is no annotated chloroplast genome available for P. fraxinifolia, which hinders the understanding of the evolution of the chloroplast genome of this species from West Asia14,15.In this study, we aim to (1) assemble and annotate the chloroplast genome of the relict species P. fraxinifolia from Hyrcanian forests; (2) perform comparative genomics of the chloroplast genomes of six Pterocarya species; and (3) assess the systematic affinity of P. fraxinifolia using phylogenetic analysis of the assembled chloroplast genomes.Materials and methodsLeaf material for the P. fraxinifolia sample was collected from a wild population in Mazandaran, Iran (Fig. 1). The voucher samples were deposited at the Herbarium of the Nowshahr Botanical Garden (HNBG) under voucher number 12,876.Fig. 1(a) Fruits in pendulous form; (b) A mature tree; (c) Regeneration under tree canopy; (d) Seedling.Full size imageGenomic DNA was extracted using the CTAB method, and its quality and quantity were checked using a Qubit 2.0 and Agilent 2100 Bioanalyzer. Libraries were created and sequenced at Wuhan Benagen Tech Solutions Company Limited, Wuhan, China, using the DNBSEQ platform (paired-end 150 bp). SOAPnuke v1.3.0 was used to filter the raw data, yielding 20 GB of clean data29.Chloroplast genome assembly and annotationRaw reads were filtered using Trimmomatic v0.3930 with a quality cutoff of 15 in a 4-base sliding window; any reads that were less than 50 bp were removed, and the adapters were filtered out. The quality of the reads before and after trimming was tested using FASTQC v0.12.1. We used GetOrganelle31 v.1.7.7.0 for chloroplast genome assembly, with the embplant_pt database used as a reference and maximum extension rounds of 15 (-R). GetOrganelle produced two isomers of the whole chloroplast genome of P. fraxinifolia, and each genome had a distinct relative orientation for the small single-copy (SSC) region32. A Python script from GetOrganelle was used along with Bowtie2 v2.5.433 to determine the average read coverage throughout the chloroplast genome. GeSeq v2.0334 was used for the initial chloroplast genome annotation of P. fraxinifolia, and the output from GeSeq was imported into Geneious Prime 2025.0.3 for an additional annotation check via the “Transfer Annotation” function. Chloroplot35 was used to produce a circular representation of the plastome.Comparative analyses of the Chloroplast genomesBecause the flanking inverted repeat (IR) regions of the chloroplast genome often vary among species, we used CPJSdraw36 to compare the IR regions of the six species. We used CUSP from EMBOSS v6.6.0.0 to calculate relative synonymous codon usage (RSCU) for protein-coding genes of P. fraxinifolia. To identify simple sequence repetitions (SSR), we used a Perl script from the Microsatellite Identification tool (MISA)37. The settings were adjusted to ten, five, and four repeats for mononucleotides, dinucleotides, and trinucleotides, respectively. Forward, reverse, palindrome, and complementary sequences with a minimum repeat length of eight bp and a maximum computed repeat of 50% were analyzed using REPuter38. The complete chloroplast genome sequences of the six Pterocarya species were aligned with Fast Statistical Alignment v1.15.938 to perform the nucleotide diversity analysis. We used a Perl script (https://github.com/xul962464/perl-Pi-nucleotide-diversity) to estimate the nucleotide diversity (PI) with a sliding window analysis with a step size of 200 bp and a window length of 800 bp.The selection pressure on chloroplast protein-coding genes (CDSs) was evaluated by aligning the nonredundant genes from six species using MAFFT v7.52639. We ran ParaAT.pl v2.040 to compute synonymous substitution rates (Ks), nonsynonymous substitution rates (Ka), and Ka/Ks. Each CDS pair of one-to-one species combinations is used as a homolog with genetic code 11. We estimated Ka, Ks, and Ka/Ks among the six Pterocarya species with KaKs_Calculator v2.041.Phylogenetic analysisWe constructed a maximum likelihood (ML) phylogenetic tree to understand the relationships of Pterocarya species. Chloroplast genome sequences were acquired from GenBank for the other Pterocarya and related genera in the Juglandaceae family. The multiple sequence alignment contained a total of 21 taxa. We performed our phylogenetic analysis using the full chloroplast genome alignment, treating it as a standard coalescent gene41. The chloroplast genomes were aligned using Fast Statistical Alignment v1.15.942 and then trimmed with trimAL v1.543 with the following settings: -automated1 -res overlap 0.7, -seqoverlap 65. To overcome the alignment issues, we also employed TAPER v1.0.047 with the -m N -a N parameters.Using RAxML-NG v1.2.144, we constructed the GTR + G model and the ML tree with 500 bootstrap repetitions. The phylogenetic tree was rooted using Engelhardia roxburghiana Wall. as an outgroup. The tree was drawn using FigTree v1.4.4 (https://github.com/rambaut/figtree). To determine the genetic distance between the six Pterocarya species, the HKY85 model45 was used, and a phylogenetic network was generated using the NeighborNet approach in SplitsTree CE v6.0.046.ResultsChloroplast genome assembly and annotationThe total numbers of raw and trimmed reads for P. fraxinifolia in this study were 143,190,876 and 141,927,817 base pairs (bp), respectively. The number of matched mapped pairs across the chloroplast genome was 393.42 ± 82.15 (Fig. S1). The complete chloroplast genome of P. fraxinifolia has a typical quadripartite structure that is 160,086 bp in length with a large single-copy region (LSC) of 89,582 bp, a small single-copy region (SSC) of 18,398 bp, and a pair of inverted repeat regions (IRs) of 26,053 bp (Fig. 2). A total of 148 genes were annotated in the chloroplast genome of P. fraxinifolia, including 103 protein-coding genes, 37 transfer RNA (tRNA) genes, and eight ribosomal RNA (rRNA) genes (Table 1 and Table S1). The GC content of the chloroplast genome was 36.17%. The annotated complete chloroplast genome of P. fraxinifolia was deposited in GenBank (accession number PV791734).Fig. 2Schematic map of overall features of the chloroplast genome of P. fraxinifolia. From the center outward, the first track shows the small single-copy (SSC), inverted repeat (IRa and IRb), and large single-copy (LSC) regions. The GC content along the genome is plotted on the second track. The genes are shown on the third track. Genes are color-coded by their functional classification. The transcription directions for the inner and outer genes are clockwise and anticlockwise, respectively. The functional classification of the genes is shown in the bottom left corner.Full size imageTable 1 Summary of the genome of Pterocarya species.Full size tableComparative analyses of the Chloroplast genome and nucleotide diversityAccording to a comparative analysis of the chloroplast genomes of Pterocarya species, the locations of eight genes in the chloroplast maps differed among species. The rps19 gene starts at position zero of the LSC region for P. fraxinifolia, but its position has shifted three times into the IRb region in the others. However, in other species of Pterocarya, a small portion of the genes were located in the IRb region. The ndhF gene in P. fraxinifolia, P. stenoptera, P. macroptera, and P. rhoifolia is located inside the SSC and is 2226 bp in length, whereas in P. tonkinensis and P. hupehensis, it spans 69 and 145 bp, respectively, into the IRb region (Fig. 3a).Fig. 3(A) Comparisons of LSC, SSC, and IR region boundaries among six Pterocarya species; (B) Nucleotide diversity (π) of CDS regions.Full size imageThe average nucleotide diversity (π) value was 0.001492, with a range of 0 to 0.00556 (Fig. 3B). The CDSs with the highest π values, which were greater than 0.0031, were ndhF, infA, ycf1, rps15, and matK. The ycf1 gene is found in the SSC area, whereas ndhF, infA, rps15, and matK are found in the LSC region. Nucleotide diversity decreased in both IR zones. Furthermore, 35 CDSs had a π value of zero among the six Pterocarya species, indicating that they were conserved (Table S1).Repeated sequence analysisThe six Pterocarya chloroplast genomes have an average of 72.6 SSR loci (Fig. 4A), with P. rhoifolia having the most SSR loci (85) of the six species (Table S2). A thorough examination of the chloroplast genome of P. fraxinifolia revealed 68 microsatellites, comprising 63 mononucleotides, four dinucleotides, and one trinucleotide simple sequence repeat. The five types of sequence repeat motifs—forward, reverse, complementary, palindromic, and tandem—are summarized in Table S3 and Fig. 4B. The analysis also revealed that the number of repetitive sequences differed across the six Pterocarya chloroplast genomes. Approximately 96.82% of the mononucleotide repeats found in P. fraxinifolia were classified as A/T (61), and 3.18% (2 repeats) were classified as C/G. In contrast, approximately 88.2% of the repeats found in P. rhoifolia were classified as A/T (75), and 3.52% (3 repeats) were classified as C/G (Fig. 4C). Dinucleotide repeats (6) for P. rhoifolia and (4) for P. fraxinifolia were the next most prevalent type of SSR. This investigation revealed no repeats of tetranucleotides, pentanucleotides, or hexanucleotides.Fig. 4Analysis of perfect simple sequence repeats (SSRs) in six Pterocarya chloroplast genomes. (A) The frequency of identified SSRs in large single-copy (LSC), inverted repeat (IR,) and small single-copy (SSC) regions; (B) The number of SSR types detected in the nine sequenced chloroplast genomes; (C) The frequency of identified SSR motifs in different repeat class types.Full size imageKa/Ks ratio and codon bias analysisStrong purifying selection and functional limitations are indicated by the very low Ka/Ks ratios found in most CDS regions among Pterocarya species (Fig. 5A). With the exception of P. tonkinensis and P. stenoptera, the highest Ka/Ks ratio was detected in the chloroplast NADH dehydrogenase F (ndhF) gene. The GC contents for the first, second, and third codon locations were 45.30%, 38.25%, and 30.36%, respectively, whereas the overall coding GC content was 37.97%. The greatest frequencies were 42.361 for the ATT codon and 37.605 for the GAA codon. The only two codons with an RSCU value of 1 were tryptophan (TGG) and methionine (ATG) (Fig. 5B). Every codon ending in A or T had an RSCU value greater than 0.5.Fig. 5Ka/Ks ratios of chloroplast protein-coding sequences across six Pterocarya species. (A) The X-axis is selected CDS with Ka/Ks ratios above 0.001. The Y-axis shows the mean Ka/Ks ratio for each gene. (B) Relative Synonymous Codon Usage (RSCU) value for each codon.Full size imagePhylogenetic analysisThe aligned multiple sequence alignment for the phylogenetic analysis consisted of 158,422 bp across 21 accessions, with 0.21% gaps and 96.19% invariant sites. The phylogenetic tree revealed Pterocarya as a sister genus to Juglans L. with 100% bootstrap support (Fig. 6A). The ML phylogenetic tree confirmed the monophyly of the genus Pterocarya with 100% bootstrap support with two subclades. P. fraxinifolia is a sister to a monophyletic subclade that include P. tonkinensis and P. macroptera and a sister to another subclade that includes P. rhoifolia, P. stenoptera, and P. hupehensis. The network analysis of the six Pterocarya species revealed a topology similar to that of the ML tree, with P. tonkinensis clustering with P. macroptera and P. rhoifolia clustering with P. stenoptera and P. hupehensis, while P. fraxinifolia branched off independently. In this study, the efficiency of two barcode regions, matK and ycf1, in the phylogeny of the genus Pterocarya was evaluated (Fig. 6B and C). The results revealed that the phylogenetic tree based on the matK region was identical to the phylogenetic tree derived from the complete chloroplast genome sequence. Pairwise distance analysis using the HKY85 method revealed that P. fraxinifolia is distantly related to Asiatic Pterocarya species (Fig. S2). The genetic distances between P. macroptera and P. tonkinensis (0.000259) and between P. stenoptera and P. hupehensis (0.000526) were the lowest, whereas the genetic distances between P. fraxinifolia and P. hupehensis (0.002153) and between P. fraxinifolia and P. tonkinensis (0.001928) were more than eightfold greater (Table S4). In the MatK dataset, P. fraxinifolia had three unique character states that differentiated it from other species of Pterocarya (Table S5).Fig. 6Comparison of three phylogenetic trees based on different chloroplast sequences: (a) Whole chloroplast genome, (b) matK gene region, and (c) ndhF gene region.Full size imageDiscussionChloroplast genomes are useful tools for studying the evolutionary relationships among species because of their preserved structure and uniparental inheritance (usually maternal in angiosperms47,48. Considering mechanisms of plant evolution49,50 and that the evolutionary history of chloroplasts is normally different from that of nuclear markers51,52, the use of genetic information from chloroplasts could reflect how seed dispersal affects the genetic makeup of wild populations and species.This study is the first to annotate the chloroplast genome of P. fraxinifolia and compare it to that of other species. We found that the positions of eight markers, namely, rps19, rpl2, ycf1 (IRa and IRb), ndhF, trnN, rpl2, and trnH, varied among the six Pterocarya chloroplast genomes. This implies that the expansion and contraction of the IR, LSC, and SSC areas are the primary sources of fluctuations in chloroplast genome size53,54. Between 68 and 85 SSRs were found among the chloroplast genomes of the six Pterocarya species. While the number of poly(G)/(C) repeats was shown to be greater in other angiosperms, the number of poly(A)/(T) repeats was significantly greater in Pterocarya.Five genes, ndhF, infA, ycf1, rps15 and matK, presented the greatest nucleotide variability (above 0.003). The matK and ycf1 genes have been suggested to function as barcode regions in plants55. The matK gene encodes the maturase protein, which facilitates the splicing of group II introns in several chloroplast genes and is considered a core barcode for land plants50,51. The ycf1 gene, which encodes the TIC214 protein that is essential for plant viability, is the second largest in the chloroplast genome and has recently been assessed for its DNA barcoding potential50,51,52, showing higher variability than the existing chloroplast candidate barcodes (such as rbcL, matK and trnH-psbA). Therefore, the ycf1 gene might be potentially useful as a DNA barcode for the Pterocarya genus56. With the exception of the matK region, none of the seven recommended barcode candidate genes in chloroplast genomes50 have the potential for barcoding of the Pterocarya genus because of a lack of nucleotide variation. Surprisingly, the accuracy of the matK region in resolving the phylogeny of the genus Pterocarya was identical to that of the complete chloroplast genome. Therefore, the matK gene alone is sufficient for reconstructing the phylogenetic relationships within the genus Pterocarya, eliminating the need for the additional time and financial resources required for whole-chloroplast-genome sequencing.The genus Pterocarya consists of six species and is closely related to Juglans in terms of pollen morphology, wood anatomy and molecular phylogenetics8,9. Our phylogenetic results confirm the sister relationship of Pterocarya to Juglans. Two sections for Pterocarya have been proposed on the basis of the presence or absence of scales on the terminal buds9,13,50. P. fraxinifolia, P. hupehensis, P. stenoptera, and P. tonkinensis belong to the section Pterocarya, while P. macroptera and P. rhoifolia belong to the section Platyptera. According to our chloroplast genome-based phylogeny, this suggested morphological classification is not supported, and the Caucasian wingnut (P. fraxinifolia) is in a distant subclade from the Chinese wingnut (P. stenoptera) and the Japanese wingnut (P. rhoifolia).The pairwise genetic distance between the Caucasian wingnut and other Asiatic Pterocarya species is greater. This distance might reflect the prolonged isolation and considerable geographic distance between Caucasian wingnut and East Asian species. Recent divergence time analyses based on fossil calibrations estimated the age of P. fraxinifolia between 9.4 and 18.4 Ma from the Miocene period and suggested the westward dispersal of Pterocarya from East Asia8. Wingnut fruit structure could facilitate the dispersal of these species by wind and water57. In this study, we collected P. fraxinifolia materials from its natural habitat in Hyrcanian forests. Our initial phylogenetic results revealed that the publicly available P. fraxinifolia in GenBank (NC046430) is not a P. fraxinifolia and is most likely a misidentified voucher that could be P. stenoptera (data not shown).Toward conservation of P. fraxinifolia
    P. fraxinifolia is classified as a vulnerable relict species on the IUCN Red List12. Our phylogenetic tree, which was constructed on the basis of chloroplast genome analysis, indicates that this species is completely distinct from other species of the genus originating from China and Japan. This distinction might highlight the species’ unique evolutionary path and specialized ecological environments. Recent studies have shown that the potentially suitable ranges of P. fraxinifolia will increase under future climate scenarios8,58, and the rapid loss of its habitat, combined with growing threats such as drought and the destruction of riparian ecosystems in Hyrcanian forests, will result in its conservation an urgent priority.

    Data availability

    The annotated complete chloroplast genome of P. fraxinifolia was deposited in GenBank, under accession number PV791734.1.
    ReferencesGrandcolas, P., Nattier, R. & Trewick, S. Relict species: a relict concept? Trends Ecol. Evol. 29, 655–663 (2014).Article 
    PubMed 

    Google Scholar 
    Ian Milne, R. Northern hemisphere plant disjunctions: A window on tertiary land bridges and climate change? Ann. Bot. 98, 465–472 (2006).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Raposo, M. & Pinto-Gomes, C. Dynamics of vegetation and climate change. Environments 9, 36 (2022).Article 

    Google Scholar 
    Zarandian, A. et al. Anthropogenic decline of ecosystem services threatens the integrity of the unique hyrcanian (Caspian) forests in Northern Iran. Forests 7, 51 (2016).Article 

    Google Scholar 
    Akhani, H., Malekmohammadi, M., Mahdavi, P., Gharibiyan, A. & Chase, M. W. Phylogenetics of the Irano-Turanian taxa of Limonium (Plumbaginaceae) based on ITS NrDNA sequences and leaf anatomy provides evidence for species delimitation and relationships of lineages: phylogenetics of IRano-TUranian Limonium. Bot. J. Linn. Soc. 171, 519–550 (2013).Article 

    Google Scholar 
    Noroozi, J. et al. Endemic diversity and distribution of the Iranian vascular flora across phytogeographical regions, biodiversity hotspots and areas of endemism. Sci. Rep. 9, 12991 (2019).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Naqinezhad, A. et al. The combined effects of climate and canopy cover changes on understorey plants of the hyrcanian forest biodiversity hotspot in Northern Iran. Glob Change Biol. 28, 1103–1118 (2022).Article 
    ADS 
    CAS 

    Google Scholar 
    Yan, H. et al. Biogeographic history of Pterocarya (Juglandaceae) inferred from phylogenomic and fossil data. J. Syst. Evol. 62, 1165–1176 (2024).Article 

    Google Scholar 
    Manning, W. E. The classification within the Juglandaceae. Ann. Mo Bot. Gard. 65, 1058 (1978).Article 

    Google Scholar 
    Usher, M. B. & Edwards, M. A dipteran from South of the Antarctic circle: belgica Antarctica (Chironomidae) with a description of its larva. Biol. J. Linn. Soc. 23, 19–31 (1984).Article 

    Google Scholar 
    Manchester, S. R. The Fossil History of the Juglandaceae (Missouri Botanical Garden, 1987).Kozlowski, G., Bétrisey, S. & Song, Y. G. Wingnuts (Pterocarya) & Walnut Family: Relict Trees: Linking the Past, Present and Future. (Natural History Museum Fribourg (NHMF), Department of Education, Culture and Sport of the State of Fribourg, Switzerland, Fribourg, (2018).Song, Y. G. et al. Phylogeny, Taxonomy, and biogeography of Pterocarya (Juglandaceae). Plants 9, 1524 (2020).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wang, T. R. et al. Adaptive divergence and genetic vulnerability of relict species under climate change: a case study of Pterocarya macroptera. Ann. Bot. 132, 241–254 (2023).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Lu, Z. J. et al. Phylogeography of Pterocarya hupehensis reveals the evolutionary patterns of a cenozoic relict tree around the Sichuan basin. For. Res. 4, e008 (2024).
    Google Scholar 
    Akhani, H. & Salimian, M. An extant disjunct stand of Pterocarya fraxinifolia (Juglandaceae) in the central Zagros Mountains, W Iran. Willdenowia 33, 113–120 (2003).Article 

    Google Scholar 
    Yao, X. et al. The first complete Chloroplast genome sequences in actinidiaceae: genome structure and comparative analysis. PLOS ONE. 10, e0129347 (2015).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Cui, G. et al. Complete Chloroplast genome of hordeum brevisubulatum: genome organization, synonymous codon usage, phylogenetic relationships, and comparative structure analysis. PLOS ONE. 16, e0261196 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Ni, Z., Zhou, P., Xin, Y., Xu, M. & Xu, L. A. Parent–offspring variation transmission in full-sib families revealed predominantly paternal inheritance of Chloroplast DNA in Pinus massoniana (Pinaceae). Tree Genet. Genomes. 17, 36 (2021).Article 
    CAS 

    Google Scholar 
    Villanueva-Corrales, S. et al. The complete Chloroplast genome of Plukenetia volubilis provides insights into the organelle inheritance. Front. Plant. Sci. 12, 667060 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Pottosin, I. & Shabala, S. Transport across Chloroplast membranes: optimizing photosynthesis for adverse environmental conditions. Mol. Plant. 9, 356–370 (2016).Article 
    CAS 
    PubMed 

    Google Scholar 
    Xia, L. et al. Chloroplast Pan-Genomes and comparative transcriptomics reveal genetic variation and temperature adaptation in the cucumber. Int. J. Mol. Sci. 24, 8943 (2023).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Li, Y. et al. New insights on the phylogeny, evolutionary history, and ecological adaptation mechanism in cycle-cup Oaks based on Chloroplast genomes. Ecol. Evol. 14, e70318 (2024).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Lim, L. W. K., Chung, H. H. & Hussain, H. Complete Chloroplast genome sequencing of Sago palm (Metroxylon Sagu Rottb.): molecular structures, comparative analysis and evolutionary significance. Gene Rep. 19, 100662 (2020).Article 
    CAS 

    Google Scholar 
    Turudić, A. et al. Variation in Chloroplast genome size: biological phenomena and technological artifacts. Plants 12, 254 (2023).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Daniell, H., Lin, C. S., Yu, M. & Chang, W. J. Chloroplast genomes: diversity, evolution, and applications in genetic engineering. Genome Biol. 17, 134 (2016).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Han, H. et al. Analysis of Chloroplast genomes provides insights into the evolution of agropyron. Front. Genet. 13, 832809 (2022).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Long, L. et al. Complete Chloroplast genomes and comparative analysis of ligustrum species. Sci. Rep. 13, 212 (2023).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Chen, Y. et al. SOAPnuke: a MapReduce acceleration-supported software for integrated quality control and preprocessing of high-throughput sequencing data. GigaScience 7, (2018).Bolger, A. et al. The genome of the stress-tolerant wild tomato species solanum pennellii. Nat. Genet. 46, 1034–1038 (2014).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Jin, J. J. et al. GetOrganelle: a fast and versatile toolkit for accurate de Novo assembly of organelle genomes. Genome Biol. 21, 241 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Palmer, J. D., Shields, C. R., Cohen, D. B. & Orton, T. J. Chloroplast DNA evolution and the origin of amphidiploid brassica species. Theor. Appl. Genet. 65, 181–189 (1983).Article 
    CAS 
    PubMed 

    Google Scholar 
    Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with bowtie 2. Nat. Methods. 9, 357–359 (2012).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Tillich, M. et al. GeSeq – versatile and accurate annotation of organelle genomes. Nucleic Acids Res. 45, W6–W11 (2017).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Zheng, S., Poczai, P., Hyvönen, J., Tang, J. & Amiryousefi, A. Chloroplot: an online program for the versatile plotting of organelle genomes. Front. Genet. 11, 576124 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Li, H. et al. CPJSdraw: analysis and visualization of junction sites of Chloroplast genomes. PeerJ 11, e15326 (2023).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Thiel, T., Michalek, W., Varshney, R. & Graner, A. Exploiting EST databases for the development and characterization of gene-derived SSR-markers in barley (Hordeum vulgare L). Theor. Appl. Genet. 106, 411–422 (2003).Article 
    CAS 
    PubMed 

    Google Scholar 
    Kurtz, S. REPuter: the manifold applications of repeat analysis on a genomic scale. Nucleic Acids Res. 29, 4633–4642 (2001).Article 
    MathSciNet 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Katoh, K. & Standley, D. M. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol. Biol. Evol. 30, 772–780 (2013).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Zhang, Z. et al. ParaAT: A parallel tool for constructing multiple protein-coding DNA alignments. Biochem. Biophys. Res. Commun. 419, 779–781 (2012).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Zhang, Z. et al. KaKs_Calculator: calculating Ka and Ks through model selection and model averaging. Genomics Proteom. Bioinf. 4, 259–263 (2006).Article 
    CAS 

    Google Scholar 
    Bradley, R. K. et al. Fast statistical alignment. PLoS Comput. Biol. 5, e1000392 (2009).Article 
    MathSciNet 
    PubMed 
    PubMed Central 

    Google Scholar 
    Capella-Gutiérrez, S., Silla-Martínez, J. M. & Gabaldón, T. TrimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25, 1972–1973 (2009).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kozlov, A. M., Darriba, D., Flouri, T., Morel, B. & Stamatakis, A. RAxML-NG: a fast, scalable and user-friendly tool for maximum likelihood phylogenetic inference. Bioinformatics 35, 4453–4455 (2019).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hasegawa, M., Kishino, H. & Yano, T. Dating of the human-ape splitting by a molecular clock of mitochondrial DNA. J. Mol. Evol. 22, 160–174 (1985).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Bryant, D. & Huson, D. H. NeighborNet: improved algorithms and implementation. Front. Bioinforma. 3, 1178600 (2023).Article 

    Google Scholar 
    Brock, J. R., Mandáková, T., McKain, M., Lysak, M. A. & Olsen, K. M. Chloroplast phylogenomics in Camelina (Brassicaceae) reveals multiple origins of polyploid species and the maternal lineage of C. sativa. Hortic. Res. 9, uhab050 (2022).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Yang, L. et al. Phylogenomic analyses reveal an allopolyploid origin of core Didymocarpinae (Gesneriaceae) followed by rapid radiation. Syst. Biol. 72, 1064–1083 (2023).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wu, F. Y., Ma, S. C., Ye, P. M., Ye, H. & Ma, J. L. The complete Chloroplast genome sequence of camellia Zhaiana (Theaceae), a critically endangered species from China. Mitochondrial DNA Part. B. 6, 2425–2426 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wei, F. et al. The complete Chloroplast genome sequence of the medicinal plant sophora tonkinensis. Sci. Rep. 10, 1–13 (2020).Article 

    Google Scholar 
    Birky, C. W. Uniparental inheritance of mitochondrial and Chloroplast genes: mechanisms and evolution. Proc. Natl. Acad. Sci. 92, 11331–11338 (1995).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Xia, M. et al. Comparative Chloroplast genome study of Zingiber in China sheds light on plastome characterization and phylogenetic relationships. Genes 15, 1484 (2024).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kim, K. J. Complete Chloroplast genome sequences from Korean ginseng (Panax schinseng Nees) and comparative analysis of sequence evolution among 17 vascular plants. DNA Res. 11, 247–261 (2004).Article 
    CAS 
    PubMed 

    Google Scholar 
    Guo, L., Zhai, J. & Gu, Y. The complete Chloroplast genome sequence of Isoetes Baodongii (Isoetaceae). Mitochondrial DNA Part. B. 9, 667–671 (2024).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Zhu, S., Liu, Q., Qiu, S., Dai, J. & Gao, X. DNA barcoding: an efficient technology to authenticate plant species of traditional Chinese medicine and recent advances. Chin. Med. 17, 1–17 (2022).Article 

    Google Scholar 
    Dong, W. et al. ycf1, the most promising plastid DNA barcode of land plants. Sci. Rep. 5, 1–5 (2015).CAS 

    Google Scholar 
    Maharramova, E. et al. Phylogeography and population genetics of the riparian relict tree Pterocarya fraxinifolia (Juglandaceae) in the South Caucasus. Syst. Biodivers. 16, 14–27 (2018).Article 

    Google Scholar 
    Song, Y., Feng, L., Alyafei, M. A. M., Jaleel, A. & Ren, M. Function of chloroplasts in plant stress responses. Int. J. Mol. Sci. 22, 13464 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Download referencesAcknowledgementsMost data analysis was performed at the Smithsonian Institution Hydra cluster https://doi.org/10.25572/SIHPC. M.V. thanks the support of Rebecca B. Dikow, Matthew Kweskin, and Eric Schuettpelz.FundingThis work was supported by a grant from the Iranian National Science Foundation (INSF), project No 4024068.Author informationAuthors and AffiliationsFaculty of Natural Sciences, Department of Environment Science, Tarbiat Modares University, Tehran, IranSeyedeh Alemeh SabbaghFaculty of Natural Sciences, Department of Forestry, Tarbiat Modares University, Tehran, IranHamed YousefzadehRoyal Botanic Gardens Kew, Richmond, Surrey, UKMohammad VatanparastDepartment of Animal and Poultry Science, College of Aburaihan, University of Tehran, Tehran, IranMohammad Reza BakhtiarizadehDepartment of Biology and Botanic Garden, University of Fribourg, Chemin du Musée 10, Fribourg, CH-1700, SwitzerlandGregor KozlowskiNatural History Museum Fribourg, Chemin du Musée 6, Fribourg, CH-1700, SwitzerlandGregor KozlowskiEastern China Conservation Centre for Wild Endangered Plant Resources, Shanghai Chenshan Botanical Garden, Shanghai, 201602, ChinaGregor Kozlowski & Yi-Gang SongAuthorsSeyedeh Alemeh SabbaghView author publicationsSearch author on:PubMed Google ScholarHamed YousefzadehView author publicationsSearch author on:PubMed Google ScholarMohammad VatanparastView author publicationsSearch author on:PubMed Google ScholarMohammad Reza BakhtiarizadehView author publicationsSearch author on:PubMed Google ScholarGregor KozlowskiView author publicationsSearch author on:PubMed Google ScholarYi-Gang SongView author publicationsSearch author on:PubMed Google ScholarContributionsH. Y. conceived and designed this study. M. V. and S. A. S conducted a formal analysis. M. B. contributed to the analytical methods. S. A. S, H. Y., and M. V. wrote the original draft. G. K. and Y. G. S. edited the manuscript. All authors have read and agreed to the published version of the manuscript.Corresponding authorsCorrespondence to
    Hamed Yousefzadeh, Mohammad Vatanparast or Yi-Gang Song.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Ethics approval and consent to participate
    The plant material of P. fraxinifolia was collected from natural populations in northern Iran under a PhD research project approved by Tarbiat Modares University, the Ministry of Science, Research and Technology of Iran. According to national regulations, the collection of plant material for academic research within Iran does not require additional permits when conducted as part of an approved university project. All sampling was done in compliance with institutional and national guidelines. We fully acknowledge the importance of adhering to the IUCN Policy Statement on Research Involving Species at Risk of Extinction as well as the Convention on the Trade in Endangered Species of Wild Fauna and Flora (CITES). We are committed to ensuring that our research complies with these guidelines and supports the conservation of endangered species.

    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 articleSabbagh, S.A., Yousefzadeh, H., Vatanparast, M. et al. Characterization of the chloroplast genome of a relict tree, Pterocarya fraxinifolia (Juglandaceae), and its comparative analysis.
    Sci Rep 15, 44153 (2025). https://doi.org/10.1038/s41598-025-23028-5Download citationReceived: 17 June 2025Accepted: 03 October 2025Published: 19 December 2025Version of record: 19 December 2025DOI: https://doi.org/10.1038/s41598-025-23028-5Share 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
    KeywordsComparative genomicsConservationHyrcanian forestsPlastome evolution More

  • in

    Coyote family activity in a landscape of fear

    AbstractCoyote (Canis latrans) presence in many North American cities evokes fear in some humans, driving demands for management action. With societal values shifting towards non-lethal coexistence practices, many wildlife managers turn to strategies like aversion conditioning, designed to increase coyotes’ fear of humans. Yet, scant knowledge exists about baseline fear behaviors (e.g., vigilance, alertness) in urban coyotes. This has implications for coexistence practices, as the motivation for coyotes’ behavior should underscore how managers respond. To explore urbanization effects on fear and behavior, we used remote cameras to monitor three coyote families during the pup-rearing season in urban, peri-urban, and rural sites in/near Calgary, Canada (2021–2022). We coded behaviors observed in adults and pups using 62 822 images. Rural adult coyotes were observed more around pups, while urban and peri-urban coyotes were observed more around pups that were playing, spent more time den-guarding, and showed higher alertness. This adaptive response in urban and peri-urban coyotes may force some coyotes into a behavioral trade-off (e.g., guarding pups vs. foraging), which could translate into more risky behaviors (e.g., consuming garbage). The elevated baseline fear in coyotes facing urban pressures suggests that coexistence practitioners must consider the risks of increasing fear during aversion conditioning.

    IntroductionGlobal urban expansion has enmeshed many wildlife species in human landscapes. Examples include coyotes in Calgary1, wild boar (Sus scrofa) in Berlin2 and caracals (Caracal caracal) in Cape Town3. It has been argued that several wildlife species now complete their entire lifecycles in cities4, leading to increased reports of human-wildlife conflict5. Coyotes are at the forefront of this conflict in North America.Evidence shows coyotes have a plethora of adaptive responses that enable urban living. For instance, coyotes can alter their population size, prey selection, and temporal use of habitat6,7,8. This resilience however can keep coyotes closer to humans, increasing the risk of negative encounters9 (e.g., a coyote bites a pet), which can cause lasting financial and emotional damage for humans and their pets.Coyote behavior throughout their lifecycle is relatively stable and predictable10, which should make coexistence straightforward. However, coexistence remains challenged by mixed human perceptions and incomplete knowledge about coyote behavior and adaptive responses in cities11. In tandem, while lethal removal of coyotes remains a core response to conflict12, there is a growing societal demand for non-lethal coexistence strategies13. This push for non-lethal approaches stems from a widespread acceptance that lethal approaches are not an effective long-term solution to human-coyote conflict, may be ecologically damaging, and are inhumane when compared to alternative non-lethal methods (e.g., aversion conditioning)14,15. Moreover, without addressing the root cause of the conflict, which is most often human behavior, there is compelling evidence that the conflict will repeat with another coyote or species15,16.In the case of urban coyotes, the pup-rearing season (May to August) is a time when conflict with humans typically spikes10 as coyotes are defensive of their pups, especially toward humans with dogs (Canis lupus familiaris)17. Previous research related to the pup-rearing stage has explored behaviors such as boldness, exploration, and aggression in both captive18 and wild coyotes19. Other work has developed a generic ethogram to measure coyote behaviors20. To our knowledge, no research has compared activity patterns and behaviors specific to pup-rearing in urban coyote families to those in less disturbed settings. Yet, changes in adaptive pup-rearing behaviors that result from urban pressure could impact conflict. For instance, a reduction in play activity amongst pups, or between pups and adults due to a heightened need for adults to be vigilant could result in less socialization of pups. In turn, lower socialization could affect pup survival after dispersal21.Consequently, we believe there exists an opportunity to develop a baseline understanding of how urbanization affects activity and behavior of coyote families. Such understanding is important in and of itself but also may foster best practices in non-lethal conflict management practice. Presently, non-lethal coexistence programs may integrate aversion conditioning (AC) to respond to conflict. Some AC approaches used on carnivores (coyotes especially) are predicated on an assumed need to re-instil or heighten an animal’s fear response to humans12,22. In the case of coyotes, some managers have used these high-intensity AC approaches when coyote behavior is labelled as ‘bold,’ ‘aggressive,’ or a risk to human safety23. Yet, no studies appear to examine fear behavior in coyotes, nor any change in behaviors because of living under urbanization pressure. If such a behavioral study demonstrated that urbanization relates to greater vigilance and fear in adult coyotes, that could be a reason to re-evaluate the use of using fear-evoking AC.To understand whether fear in coyotes is affected by urbanization we must first address the question “What is natural fear in wild animals?” Here, we can borrow from the “landscape of fear” concept, in which prey species will adjust their behavior in response to the threat of predation24. This concept derives from fear ecology work in which the behavioral response of prey was compared between fear-driven systems and direct predation-driven systems25. The concept explains how the knowledge of predation risk impacts prey choices in space use, foraging, and vigilance behavior26. For example, black-tailed jackrabbits (Lepus californicus) and desert cottontails (Sylvilagus audobonii) adjust their movements and behaviors in response to fear of predators27. The landscape of fear concept also has been invoked in situations where humans play the role of the “super-predator,” such as with elk (Cervus canadensis)28 and marsupials29. Importantly though, while fear can be adaptive for reducing predation risk and increasing life expectancy in prey species30, it can come at a cost of foraging31. In turn, this can add food stress, which can increase conflict within the prey populations. Similarly, human activity may drive changes in top predator behavior, impacting the animals’ ability to regulate prey populations. In the case of coyotes – the top predator in many cities – additive fear in a landscape could lead to unchecked populations of small mammal species32 and larger mammals like deer (Odocoileus spp)32. This may lead to spill-over effects for humans, such as rodent infestations or greater deer-vehicle impacts33. Most critically, if coyotes spend more time in behaviors that arise out of fear (e.g., vigilance) at the expense of foraging, this could lead to opportunistic feeding on anthropogenic food sources, which may increase stress and conflict in coyotes34. Fear in urban-adapted wildlife has been previously studied in the foraging behavior of smaller, prey mammals, with lower levels of vigilance in treatments closer to urban areas but higher responses to fear stimuli in treatments in a peri-urban environment35. However, fear in non-foraging coyotes across an urban to rural gradient is not well understood.To explore whether urbanization affected adult coyote activity and behavior during the pup-rearing season, and specifically whether fear was higher in urban coyotes, we narrowed our analysis to the following general questions: (1) Were there changes in adult presence and fear-related behavior (i.e., den-guarding) around pups across rural to urban sites?; (2) Did the percent of time spent by adult coyotes on high alert during captured activity sequences increase across the rural to urban gradient?ResultsWe coded 81 442 images from camera traps (CTs), of which 62 822 captures showed coyotes (including adults and pups), across 923 total trap nights (402 from Campus [the urban site], 188 from Spyhill [the peri-urban site], and 333 from WA Ranches [the rural site]). We used a selection of the coded CT photos that were isolated to the pup-rearing season in 2022, totalling 15 000 captures from Campus, 15 108 captures from Spyhill, and 14 808 captures from WA Ranches. Amongst the latter subset, 21 386 contained adults, 31 663 contained pups, and 8 513 had both pups and adults present. We converted the CT photos into 4 556 sequences of activity across the entire sample area, 2 021 of which were used for proportional analysis of behavior Fig 1.Fig. 1Study site locations within and around the city of Calgary, Alberta. The arrow indicates the direction of the gradient of urban to rural.Full size imageAdult Coyote presence and behavior around pupsIn our first set of comparisons, we examined adult behavior and presence around pups and pup play (Fig. 2). We observed adult attendance to pups in behaviors such as interacting, nursing, and guarding (Table 1). While pups were attended by adults in ~ 27% of photos across all three sites, we found significant differences between adult presence around pups by site (X2 = 37.717, P = 6.454e-09, df = 2), adult presence around pups playing by site (X2 = 137.63, P < 2.2e-16, df = 2), and den guarding around pups by site (X2 = 352.2, P < 2.2e-16, df = 2). Using Pearson residuals, we investigated which frequencies deviated the most from what would be expected if there was no difference between sites. We found significantly more adult presence around pups at WA Ranches (i.e., rural). We observed significantly greater adult presence around pups that were playing and adult guarding at Campus (i.e., urban).Fig. 2Proportions of photos of adult presence and behavior around pups. Proportions were determined as the number of photos withing a subset of photos of pups or of pups playing that display the behavior of interest (i.e. guarding) over the total number of photos within the subset. Stars indicate significant contributions to deviance from independence, as determined by Pearson’s residuals (P < 0.05).Full size imageTable 1 Description of all Coyote behaviors captured. Assessment of behaviors was based on previous behavioral studies20,55,56.Full size tableAdult Coyote alertness by site and conditional variablesWe compared the percent of activity time spent by adults on high alert by site including other conditional variables (e.g., pup presence, time of day, novel object presence). The mean proportion of images showing high alert behavior relative to not high alert behavior was 12.8% across all three sites during the spring/summer sample period. At Campus, the mean proportion of high alert behavior was 16.7%, at Spyhill it was 14.3%, and at WA Ranches it was 4.8%. The results of our zero-inflated binomial mixed-effects model of the percentage of high alert behavior per activity sequence, including all independent variables with significant interactions with each other, are presented in Table 2. Note that for novel object presence, only six out of 2 021 image sequences captured a novel object.Table 2 Results of the zero-inflated binomial mixed-effects model on the proportion of high alert behavior by sequence. Data come from the May-August 2022 sample period (n = 2 021).Full size tableDue to the high number of pairwise interactions between independent variables in the model, the post-hoc Tukey analysis was Sidak-adjusted for the comparison of means36. The variation in these means of the proportions of high alert behavior per sequence by site, time of day, and pup presence is presented in Fig. 3. The highest estimated marginal means for the proportion of high alert behavior per sequence occurred at the Campus site in the daytime with pups present, while the lowest occurrence of high alert behavior occurred at the WA Ranches site at twilight with pups present. Significant differences included higher marginal means for alertness at Campus than WA Ranches at twilight without pups present (Z ratio = 2.638, P = 0.0227, df = inf) and at any time of day with pups present (Day: Z ratio = 4.849, P < 0.0001, df = inf; Night: Z ratio = 3.036, P = 0.0068, df = inf; Twilight: Z ratio = 4.499, P < 0.0001, df = inf), and higher marginal means for alertness at Spyhill than WA Ranches in the daytime with pups present (Z ratio = 2.473, P < 0.0357, df = inf) and at twilight with pups present (Z ratio = 3.182, P < 0.0042, df = inf). Degrees of freedom are labeled as infinite using the emmeans package36 as estimates were compared against the standard normal distribution. Fig. 3Pairwise comparison of marginal means from the zero-inflated binomial mixed effects model on the proportion of high alert behavior per sequence. The interaction term effects are shown for the relationship between adult coyote vigilance and study site, pup presence, and time of day. Boxes indicate the marginal mean while the error bars indicate the 95% confidence interval of the marginal mean. The marginal means were determined using the emmeans package36 while the visual was created using the ggplot2 package63.Full size imageDiscussionCoyotes have a high investment in their pups, as seen by the frequent attendance of them by both parents37 and the contribution of non-breeding helpers to pup-rearing38. At the WA Ranches (i.e., rural) site, we observed slightly more adult presence with the pups, which could be a positive indicator for pup survival, as reported previously by Bekoff and Wells38. However, when engaged in play activities, we observed that the pups were left unsupervised significantly more at WA Ranches and Spyhill (i.e., peri-urban) when compared to Campus (i.e., urban). We also observed that guarding behavior by adults was significantly more common amongst the Campus coyotes. One explanation is that the Campus coyote family perceived this urban homesite to be riskier for pups due to being embedded in high density urban matrix. When pups engaged in play, we observed routinely that they spread quickly outside the visual range of adult coyotes. In an urban site, the higher incidence of potentially dangerous novel objects may demand higher levels of pup supervision and guarding by adults.The significantly higher incidence of guarding pups at Campus could indicate that urban coyotes may experience more concern and fear for themselves and their offspring. To mitigate the risks to pups, one might expect that adult coyotes may have to adjust their activity budgets. Particularly concerning would be if increased time spent in pup-supervision leads to adult coyotes making trade-offs with essential behaviors like hunting. The latter was observed with African wild dogs (Lycaon pictus), where an increase in pup-guarding, though effective in decreasing pup mortality, negatively impacted hunting as the wild dog groups were constrained by the absence of pup-guarders to assist with hunts39. Arguably, we are seeing evidence that suggests urban coyotes (at least in the Campus focal family) may be redirecting time to protect pups.Another potential indicator of a heightened fear response is seen in vigilance behavior. We measured vigilance as high alert behavior and den-guarding. Vigilance has been associated with fear hormonally as it can be paired with the release of fear-response hormones40. Vigilance has been observed to increase in other species around human presence28,32,41, but it has not been studied before in this manner in wild coyotes. In previous studies, vigilance has been measured in behavioral experiments of urban coyotes19 and in understanding the response of captive coyotes to human activity42. We found that adult coyotes spent significantly more time displaying vigilance behaviors (high alertness, den-guarding) at the urban site compared to the peri-urban and rural sites. For instance, in the rural site (WA Ranches), where coyotes share the landscape with larger predators caught on camera like grizzly bears (Ursus arctos) and cougars (Puma concolor), rural coyotes only spent ~ 5% of their time on high alert, while the urban coyotes spent ~ 17% of their time on high alert (as captured in the data). Given the potential for a trade-off to occur between fear and foraging31, the significant increase in time spent in vigilance (assuming it is fear related) by urban coyotes may drive them to procure opportunistic easier anthropogenic food sources like garbage, which can increase human-coyote conflict34,43. This behavioral shift may in turn be paired with an avoidance of certain high-human use areas, potentially resulting in an increase in prey animal presence, as has been suggested in the “human-shield hypothesis”45.In grouping all three coyote families, we also found that alertness was higher overall in the presence of pups. This would be expected, given demonstrated parental investment by coyote parents in their young37. However, examining WA Ranches on its own, adult alertness was significantly lower in the presence of pups than the absence of pups. This may suggest that these rural coyotes perceive less risk and therefore experience a lower baseline level of fear than their urban cousins. Certainly, the rural site is protected from hunting, trapping, and poisoning (within the ranch) and very few people enter the site, which would create a sense of security from humans.When examining the relationship between time of day and high alertness, we found a significant increase in alertness during the twilight period. We observed more high alertness overall in the twilight period with no significant variation between sites but less high alertness in the presence of pups in the twilight and night periods, suggesting there is heightened concern for pups in the daylight. Previous urban coyote studies have shown a decrease in coyote activity in the twilight period corresponding with higher human activity, suggesting coyotes use an avoidance technique at this time of day45. As the coyotes in this study were viewed at homesites where they can avoid humans, the heightened alertness may be another response to an increase in human activity.Examining the relationship between the length of an activity sequence (as determined by the number of images of coyotes in the sequence) and high alertness, there was a slight but significant negative effect of more photos. This may relate to longer sequences having more behaviors visible, which could offset the overall presence of high alertness; alternatively, a short sequence of photos could have three photos all of which display high alertness making for a much higher proportion.Exploring the relationship between the maximum number of adults in a sequence and high alertness, there was a significant decrease in high alertness with more adults. More adults in the sequence meant a lower proportion of time spent on high alert which corroborates the hypothesis of the negative relationship between vigilance and group size seen among other species46.Finally, in our investigation of how the presence of a novel object related to high alertness, we found a significant negative effect. That said, we rarely captured novel objects in the photos (only six out of 2 021 sequences featured novel objects) and this result may have occurred out of chance as the few observations happened to occur when coyotes were on lower alert. Novel objects in the study were sporadically occurring objects observed to have drifted into view of the cameras. Novel objects also typically occurred in the urban site where coyotes may be more adjusted to the presence of such things and have less of a response. Urban coyotes have also been observed to be more investigative in the presence of novel objects19, so this data could simply be a further indication of this behavior. This will be an important area of future research to understand whether the likelihood of encountering many more novel objects in an urban ecosystem means that adult coyotes spend even more time supervising pups at the expense of other vital behaviors.Our study provides evidence for a heightened state of fear-related behavior among urban relative to peri-urban and rural coyote families. As noted, this can come at a cost of adaptive behaviors, as seen in other species32,39. Because fear is associated with stress and stress can lead to riskier behavior in coyotes and conflict with humans23, understanding that urban coyotes exhibit significantly more fear daily is important to coexistence practices. In particular, non-lethal coyote management strategies that implement fear-based AC methods have not to our knowledge evaluated what baseline levels of fear exist, or whether adding stress or fear may have compounding negative impacts on coexistence. In our opinion, coexistence programs should consider the efficacy and ethics of programs that ‘stack’ fear upon fear47. We observed significantly higher rates of vigilance behavior related to pup guarding, which suggests coyotes may be more fearful during that time. Therefore, there is a risk that using fear-based AC with urban coyotes may create a condition called “trigger stacking”48,49. In domestic dogs, trigger stacking is known to exacerbate reactivity rather than change or de-escalate behavior48. If coyotes might become more reactive due to AC-caused trigger stacking that would be counterintuitive to the goals of coexistence.While urban coyotes have been characterized as bolder and more exploratory, our results highlight fearfulness (i.e., guarding, pup-attendance, vigilance) as another key element motivating their behavior. If more fear leads to restricting mobility and access to natural foods, this could drive consumption of more easily accessible anthropogenic food, which has historically increased human-coyote conflict. While we examined only three families, one per level of urbanization, the magnitude of difference in behaviors between the sites highlights the need to better understand the baseline ecology and behavioral adaptations of coyotes before applying untested invasive coexistence management techniques. Further study could benefit from exploring behaviors outside of the pup-rearing season and look at direct impacts of such management on coyote and other urban wildlife behaviors.MethodsStudy areaOur research was conducted at three sites in and around the City of Calgary, Alberta, Canada (Fig. 1). The study sites rest within the Foothills Parkland Natural Subregions, a hilly area with a mixture of grasslands, shrublands, and forest that lies between the prairies to the east and the foothills to the west49. The region is home to a diverse array of plant species and several mammal species from hares (Lepus spp.) to coyotes to bears (Ursus spp.). The region has a relatively dry and cold climate, with a mean temperature of 4.3 °C ranging from − 30 °C to + 30 °C and a mean precipitation of 417 mm of rain and 100 cm of snow50. The research reported was covered by Animal Care Certificate number AC20-0160 issued by the University of Calgary Life and Environment Sciences Animal Care Committee for the project “UC Campus Coyote Ecology and Coexistence” on February 23, 2021. The experiments reported in this manuscript were minimally invasive and conducted in accordance with Animal Care Committee guidelines. Clinical trial number: not applicable.Each study site represented a unique level of human use, with the Campus site (i.e., urban) located within the City of Calgary and surrounded by residential area, the Spyhill site (i.e., peri-urban) located on the northwest edge of the city, at the juncture of urban residential and agricultural land use, and the WA Ranches (i.e., rural) site located approximately 30 km northwest of Calgary, surrounded by ranch land and natural areas. Each site was the core area of a unique coyote family comprised of a breeding pair, one ‘helper,’ and an annual litter of three to eight pups (SM Alexander, unpublished data). Sites were approximately 30 km apart, which is outside the limits of a resident home range51 and reduced the chances of detecting the same coyote at different sites. Our visual records showed that there was no observed overlap of individuals from different families, even though as Gehrt et al.53 note, this distance can easily allow transient coyotes to cross amongst sites. AC by agencies was only reported to have occurred within the core habitat of Campus coyotes.Camera-Trap methodsWe deployed 27 camera traps (CTs), divided equally by site (21 Reconyx Hyperfire 2 and 6 Cabela’s Outfitter Gen 3). Each camera captured three images at a detected motion, with images continuously captured if motion continued, resulting in a sequence of activity. Images were captured at a rate of one image per second. The resolution of images was 16 megapixels for the Cabela’s cameras and two megapixels for the Reconyx cameras. No differences were noted in the ability of either camera to capture coyote activity, though both would occasionally cease functioning in extremely cold temperatures (i.e., less than − 20 °C). Cameras were set to operate at all times of day and night, only stopping if the batteries died or the SD card storage was filled, but the frequency of camera checks allowed them to run for the most part continuously. The time and date were set on the camera at its placement. Cameras were placed on trees or fences at 30–60 cm off the ground to maximize the field of view for capturing coyotes. Instead of using bait, the cameras were pointed toward known high-coyote-activity areas, such as around the home sites and high-use pathways, as determined from field surveying. The camera locations were purposefully selected as the goal was to capture the highest amount of coyote activity. While a random selection may have captured a more natural range of coyote behaviors throughout their territory, the focus here was specific behaviors at high-use areas within the homesite and comparing these behaviors between sites.All cameras were in place year-round as part of long-term monitoring of coyotes, but for this project we screened focal images from CTs at post-natal homesites only for the period of May/June 2021 and January to August 2022 from Campus (48 290 images) and May to August 2022 from Spyhill (15 893 images) and WA Ranches (17 259). We focused on the pup-rearing season to capture and compare behavior and activity budgets when we were most likely to see interactions amongst coyote family members at the three sites of interest. We divided the CT photos of coyotes into sequences of activity, using a separation of one minute between a coyote disappearing and reappearing on the screen. Within these sequences we could then determine the proportion of time spent displaying any one behavior (i.e., calculated as the number of captures displaying the behavior over the total captures in the sequence). While we used shorter image capture intervals than other studies, such as 5 min in Wooster et al.54 and 10 min in Marion et al.42. Our method suited the resolution of our research question; We had no need to try to separate sequences by unique individuals, as known individuals frequented our same site and our exceptionally large photo counts increased replicates when compared to other noted studies.We developed an ethogram in reference to our CT data and classified fear-related behaviors for coyotes as alertness and pup-rearing behaviors (e.g., den-guarding). The ethogram was founded on previous behavioral research on coyotes54, red foxes55, and felids56 and honed to the study animals during initial reviews of the photos. We describe all behaviors documented and how the behaviors were coded in Table 1. For each CT photo sequence of activity, we documented the following: proportion of time spent in the behavior, site, time of day, pup presence, number of photos, maximum number of adults, novel object presence, date, and camera location. One observer classified all photos into their behavioral categories.Statistical analysesTo determine whether coyote behavior around pups varied by category (type) amongst study sites, we performed multiple chi-squared tests57. For this analysis, we only used images that showed pups to be present (i.e., if photos were adults only, we removed them from this analysis). We compared the frequency of the photos showing the following behaviors across sites: pup photos with and without adult presence, pup play photos with and without adult presence, and pup photos with adults demonstrating a den-guarding posture or not. We used Pearson residuals to determine which behaviors differed significantly by site, visualizing the differences using the vcd package58.To explore whether fear-related behaviors differed across human disturbance categories, we developed a generalized linear model and compared the proportion of photos per sequence demonstrating high alertness. We included variables previously identified to be relevant to behavior: site, pup presence, time of day, sequence length, maximum adult presence, and novel object presence37. We also were interested in how pup presence and time of day might interact with site, so we included an interaction term for those three independent variables. Data were proportional and thus followed a binomial distribution57. Given many sequences of activity had proportions of zero for high alert behavior, we used a zero-inflated model using the glmmTMB package59. Since our data included multiple images coming from the same cameras and from each site, we followed Zuur and Ieno60, using a mixed effects model with a random effect of camera nested within site to account for this level of variation. Following from the previous, our final model was a zero-inflated binomial mixed effects model, described as:proportion of high alert behavior per sequence ~ site*time of day*pup presence + number of photos in the sequence + maximum number of adults in the sequence + novel object presence in the sequence.We performed a post-hoc Tukey analysis of multiple comparisons61 using emmeans package36 to identify pair-wise interactions between model terms. All statistical analyses were performed using R version 3.4.263.

    Data availability

    Data is provided within the manuscript or supplementary information files.
    ReferencesLukasik, V. M. & Alexander, S. M. Human–Coyote interactions in Calgary, Alberta. Hum. Dimensions Wildl. 16, 114–127 (2011).Article 

    Google Scholar 
    Stillfried, M. et al. Secrets of success in a landscape of fear: urban wild Boar adjust risk perception and tolerate disturbance. Front. Ecol. Evol. 5, 157 (2017).Article 

    Google Scholar 
    Nattrass, N. & O’riain, M. J. Contested natures: conflict over caracals and cats in cape Town, South Africa. J. Urban Ecol. 6, 1–12 (2020).Orthmeyer, D. L., Cox, T. A., Turman, J. W. & Bennett, J. R. Operational Challenges of Solving Urban Coyote Problems in Southern California. Wildlife Damage Management Conference Proceedings https://digitalcommons.unl.edu/icwdm_wdmconfproc/69 (2007).Seto, K. C., Güneralp, B. & Hutyra, L. R. Global forecasts of urban expansion to 2030 and direct impacts on biodiversity and carbon pools. Proc. Natl. Acad. Sci. U S A. 109, 16083–16088 (2012).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Elliot, E. E., Vallance, S. & Molles, L. E. Coexisting with Coyotes (Canis latrans) in an urban environment. Urban Ecosyst. 19, 1335–1350 (2016).Article 

    Google Scholar 
    Mitchell, N., Strohbach, M. W., Pratt, R., Finn, W. C. & Strauss, E. G. Space use by resident and transient Coyotes in an urban–rural landscape mosaic. Wildl. Res. 42, 461–469 (2015).Article 

    Google Scholar 
    White, L. A. & Gehrt, S. D. Coyote attacks on humans in the united States and Canada. Hum. Dimensions Wildl. 14, 419–432 (2009).Article 

    Google Scholar 
    Baker, R. O. & Timm, R. M. Coyote attacks on humans, 1970–2015: implications for reducing the risks. Human-Wildlife Interact. 11, 120–132 (2017).
    Google Scholar 
    Bekoff, M. & Gese, E. M. Coyote (Canis latrans). USDA National Wildlife Research Center – Staff Publications 224, Fort Collins, Colorado, USA https://digitalcommons.unl.edu/icwdm_usdanwrc/224 (2003).Drake, M. D. et al. How urban identity, affect, and knowledge predict perceptions about coyotes and their management. Anthrozoos 33, 5–19 (2020).Plotsky, K., Alexander, S. M., Musiani, M. & Draper, D. Incorporating geographic context into Coyote and Wolf livestock depredation research. Can. Geogr. / Le Géographe Canadien. 66, 450–461 (2022).
    Google Scholar 
    Hunold, C. & Mazuchowski, M. Human–wildlife coexistence in urban wildlife management: insights from nonlethal predator management and rodenticide bans. Animals 10, 1983 (2020).McManus, J. S., Dickman, A. J., Gaynor, D., Smuts, B. H. & Macdonald, D. W. Dead or alive? Comparing costs and benefits of lethal and non-lethal human–wildlife conflict mitigation on livestock farms. Oryx 49, 687–695 (2015).Article 

    Google Scholar 
    Treves, A., Krofel, M. & McManus, J. Predator control should not be a shot in the dark. Front. Ecol. Environ. 14, 380–388 (2016).Article 

    Google Scholar 
    Fox, C. H. & Papouchis, C. M. Coyotes in our midst: Coexisting with an adaptable and resilient carnivore. Animal Protection Institute. Sacramento, California, USA. http://www.projectcoyote.org/events/nov302010event.pdf (2005).Quinn, N., Fox, D., Hartman, J. & Org, E. An Examination of Citizen-Provided Coyote Reports: Temporal and Spatial Patterns and Their Implications for Management of Human-Coyote Conflicts. Proceedings of the Vertebrate Pest Conference 27, 27 (2016).Young, J. K., Mahe, M. & Breck, S. Evaluating behavioral syndromes in Coyotes (Canis latrans). J. Ethol. 33, 137–144 (2015).Article 

    Google Scholar 
    Breck, S. W., Poessel, S. A., Mahoney, P. & Young, J. K. The intrepid urban coyote: a comparison of bold and exploratory behavior in Coyotes from urban and rural environments. Sci. Rep. 9, 1–11 (2019).Article 
    CAS 

    Google Scholar 
    Way, J. G., Szumylo, D. L. M. & Strauss, E. G. An ethogram developed on captive Eastern Coyotes canis latrans. Can. Field-Naturalist. 120, 263–288 (2006).Article 

    Google Scholar 
    Bekoff, M. & Wells, M. C. Social ecology and behavior of Coyotes. Adv. Study Behav. 16, 251–338 (1986).Article 

    Google Scholar 
    Breck, S. W., Poessel, S. A. & Bonnell, M. A. Evaluating lethal and nonlethal management options for urban Coyotes. Human-Wildlife Interact. 11, 133–145 (2017).
    Google Scholar 
    Timm, R. M., Baker, R. O., Bennett, J. R. & Coolahan, C. C. Coyote attacks: an increasing suburban problem. Trans. North. Am. Wildl. Nat. Resour. Conf. 69, 67–88 (2004).
    Google Scholar 
    Laundré, J. W., Hernández, L. & Altendorf, K. B. Wolves, elk, and bison: reestablishing the ‘landscape of fear’ in Yellowstone National Park, U.S.A. Can. J. Zool. 79, 1401–1409 (2001).Article 

    Google Scholar 
    Brown, J. S., Laundré, J. W. & Gurung, M. The ecology of fear: optimal foraging, game theory, and trophic interactions. J. Mammal. 80, 385–399 (1999).Article 

    Google Scholar 
    Laundre, J. W., Hernandez, L. & Ripple, W. J. The landscape of fear: ecological implications of being afraid. Open. Ecol. J. 3, 1–7 (2010).Article 

    Google Scholar 
    Razo, A. D., Hernández, I., Laundré, L., Velasco-Vázquez, L. & J. W. & The landscape of fear: habitat use by a predator (Canis latrans) and its main prey (Lepus Californicus and Sylvilagus audubonii). Can. J. Zool. 90, 683–693 (2012).Article 

    Google Scholar 
    Ciuti, S. et al. Effects of humans on behaviour of wildlife exceed those of natural predators in a landscape of fear. PLoS One. 7, e50611 (2012).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    McGann, K., Johnson, C. N., Clinchy, M., Zanette, L. Y. & Cunningham, C. X. Fear of the human ‘super predator’ in native marsupials and introduced deer in Australia. Proc. Biol. Sci. 291, 20232849 (2024).PubMed 
    PubMed Central 

    Google Scholar 
    Boissy, A. Fear and fearfulness in animals. Q. Rev. Biology. 70, 165–191 (1995).Article 
    CAS 

    Google Scholar 
    Brown, J. S. & Kotler, B. P. Hazardous duty pay and the foraging cost of predation. Ecol. Lett. 7, 999–1014 (2004).Article 

    Google Scholar 
    Suraci, J. P., Clinchy, M., Zanette, L. Y. & Wilmers, C. C. Fear of humans as apex predators has landscape-scale impacts from mountain lions to mice. Ecol. Lett. 22, 1578–1586 (2019).Article 
    PubMed 

    Google Scholar 
    Gilbert, S. L. et al. Socioeconomic benefits of large carnivore recolonization through reduced Wildlife-Vehicle collisions. Conserv. Lett. 10, 431–439 (2017).Article 

    Google Scholar 
    Murray, M. H. & St. Clair, C. C. Individual flexibility in nocturnal activity reduces risk of road mortality for an urban carnivore. Behavioral Ecology 26, 1520–1527 (2015).Fardell, L. L., Nano, C. E. M., Pavey, C. R. & Dickman, C. R. Small prey animal foraging behaviors in landscapes of fear: effects of predator presence and human activity along an urban disturbance gradient. Front. Ecol. Evol. 10, 805891 (2022).Article 

    Google Scholar 
    Lenth, R. V. et al. emmeans: estimated marginal means, aka least-squares means. Preprint at (2023).Harrison, D. J. & Gilbert, J. R. Denning ecology and movements of Coyotes in Maine during pup rearing. J. Mammal. 66, 712–719 (1985).Article 

    Google Scholar 
    Bekoff, M., Wells, M. C. & Defense, R. Behavioral Ecology of Coyotes: Social Organization, Rearing Patterns, Space Use, and Z Tierpsychol 60, 281–305 (1982).Courchamp, F., Rasmussen, G. S. A. & Macdonald, D. W. Small pack size imposes a trade-off between hunting and pup-guarding in the painted hunting dog Lycaon pictus. Behav. Ecol. 13, 20–27 (2002).Article 

    Google Scholar 
    McCarty, R. & Kopin, I. J. Changes in plasma catecholamines and behavior of rats during the anticipation of footshock. Horm. Behav. 11, 248–257 (1978).Article 
    CAS 
    PubMed 

    Google Scholar 
    Marion, S. et al. Red deer behavioural response to hiking activity: a study using camera traps. J. Zool. 317, 249–261 (2022).Article 

    Google Scholar 
    Schultz, J. T. & Young, J. K. Behavioral and Spatial responses of captive Coyotes to human activity. Appl. Anim. Behav. Sci. 205, 83–88 (2018).Article 

    Google Scholar 
    Alexander, S. M., Quinn, M. S. & Coyote Canis latrans) interactions with humans and pets reported in the Canadian print media (1995–2010). Hum. Dimensions Wildl. 16, 345–359 (2011).Article 

    Google Scholar 
    Gaynor, K. M. et al. The human shield hypothesis: does predator avoidance of humans create refuges for prey? Ecol. Lett. 28, e70138 (2025).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Gese, E. M., Morey, P. S. & Gehrt, S. D. Influence of the urban matrix on space use of Coyotes in the Chicago metropolitan area. J. Ethol. 30, 413–425 (2012).Article 

    Google Scholar 
    Elgar, M. A. Predator vigilance and group size in mammals and birds: a critical review of the empirical evidence. Biol. Reviews – Camb. Philosophical Soc. 64, 13–33 (1989).Article 
    CAS 

    Google Scholar 
    Shreyer, T., Barrett, S. & Shull, A. Addressing Canine and Feline Behavior Problems in Clinical Practice. in Introduction to Animal Behavior and Veterinary Behavioral Medicine 243–268 (John Wiley & Sons, (2024).Edwards, P. T., Smith, B. P., McArthur, M. L. & Hazel, S. J. Fearful fido: investigating dog experience in the veterinary context in an effort to reduce distress. Appl. Anim. Behav. Sci. 213, 14–25 (2019).Article 

    Google Scholar 
    Government of Alberta. Natural Regions Committee. Natural Regions and Subregions of Alberta. http://www.cd.gov.ab.ca/preserving/parks/anhic/Natural_region_report.asphttps://doi.org/10.5962/bhl.title.115367 (2016).The City of Calgary. Climate projections for Calgary. (2022). https://regionaldashboard.alberta.ca/region/calgary/#/Grinder, M. I., Krausman, P. R., Home & Range Habitat Use, and nocturnal activity of Coyotes in an urban environment. J. Wildl. Manage. 65, 887 (2001).Article 

    Google Scholar 
    Gehrt, S. D., Anchor, C. & White, L. A. Home range and landscape use of Coyotes in a metropolitan landscape: conflict or coexistence? J. Mammal. 90, 1045–1057 (2009).Article 

    Google Scholar 
    Wooster, E., Wallach, A. D. & Ramp, D. The wily and courageous red fox: behavioural analysis of a mesopredator at resource points shared by an apex predator. Anim. 2019. 9, 907 (2019).
    Google Scholar 
    Way, J. G., Auger, P. J., Ortega, I. M. & Strauss, E. G. Eastern Coyote Denning behavior in an anthropogenic environment. Northeast Wildl. 56, 18–30 (2001).
    Google Scholar 
    Wooster, E. I. F., Ramp, D., Lundgren, E. J., O’Neill, A. J. & Wallach, A. D. Red foxes avoid apex predation without increasing fear. Behav. Ecol. 32, 895–902 (2021).Article 

    Google Scholar 
    Stanton, L. A., Sullivan, M. S. & Fazio, J. M. A standardized ethogram for the felidae: A tool for behavioral researchers. Appl. Anim. Behav. Sci. 173, 3–16 (2015).Article 

    Google Scholar 
    Motulsky, H. Intuitive Biostatistics: A Nonmathematical Guide To Statistical Thinking (Oxford University Press, 2014).Zeileis, A., Meyer, D. & Hornik, K. Residual-Based shadings for visualizing (Conditional) independence. J. Comput. Graphical Stat. 16, 507–525 (2007).Article 
    MathSciNet 

    Google Scholar 
    Brooks, M. E. et al. GlmmTMB balances speed and flexibility among packages for zero-inflated generalized linear mixed modeling. R J. 9, 378–400 (2017).Article 

    Google Scholar 
    Zuur, A. F. & Ieno, E. N. A protocol for conducting and presenting results of regression-type analyses. Methods Ecol. Evol. 7, 636–645 (2016).Article 

    Google Scholar 
    Tukey, J. W. Comparing individual means in the analysis of variance. Biometrics 5, 99 (1949).Article 
    MathSciNet 
    CAS 
    PubMed 

    Google Scholar 
    R Core Team. R: a language and environment for statistical computing. Preprint at. (2022).Wickham, H. Ggplot2: Elegant Graphics for Data Analysis. (2016).Download referencesAcknowledgementsThis research was funded in part by the Natural Sciences and Engineering Research Council of Canada, the Social Science and Humanities Research Council, and the University of Calgary, Canada.Author informationAuthors and AffiliationsUniversity of Calgary, Calgary, AB, CanadaRobert Mitchell & Shelley AlexanderAuthorsRobert MitchellView author publicationsSearch author on:PubMed Google ScholarShelley AlexanderView author publicationsSearch author on:PubMed Google ScholarContributionsR.M. and S.A. conceived the experiment, R.M. and S.A. conducted the experiment, R.M. analysed the results. All authors reviewed the manuscript.Corresponding authorCorrespondence to
    Shelley Alexander.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 articleMitchell, R., Alexander, S. Coyote family activity in a landscape of fear.
    Sci Rep 15, 44210 (2025). https://doi.org/10.1038/s41598-025-28363-1Download citationReceived: 24 March 2025Accepted: 10 November 2025Published: 19 December 2025Version of record: 19 December 2025DOI: https://doi.org/10.1038/s41598-025-28363-1Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Short- and long-term effects of culling invasive corallivorous gastropods

    AbstractEradicating invasive species and maintaining their populations at acceptable densities is both costly and challenging in marine environments, primarily due to the open water connectivity between culled and non-culled areas. This research aims to evaluate the short- and long-term effects of culling invasive species, considering the invasive gastropod Drupella rugosa (Born, 1778) from the coral reefs of Koh Tao (Gulf of Thailand) as a case study. Ecological, logistical, and behavioural factors that influenced the removal efforts were identified, highlighting key components that can inform future strategies aimed at managing outbreak events. Specific objectives included: (1) estimating gastropod densities and to study the behaviour of D. rugosa on Acropora-dominated reefs; (2) assessing short-term effects of D. rugosa removal by monitoring the fate of grazed corals; (3) examining the long-term impact of culling by analysing data from a removal campaign spanning over a decade, including an evaluation of the effort in terms of time and diver involvement. The relationship between damselfish and the feeding activity of corallivorous gastropods was also investigated. A key finding of this study is that poorly planned culling is ineffective in controlling outbreaks of invasive species such as those belonging to the genus Drupella. Long-term data from culling campaigns conducted between 2010 and 2024 revealed that the number of removed specimens remained relatively constant, despite significant differences in effort. This disparity underscores the lack of strategic coordination in the implementation of removal activities. Following a critical comparison with cases reported in the literature, common issues and transferable strategies were identified and thoroughly analyzed. Directions for management were provided, with the understanding that future actions should be grounded in a thorough knowledge of the species’ ecological traits, the biotic and abiotic drivers of outbreak events, a quantitative assessment of its impact on Acropora reefs, and integration into with well-established international removal and prevention programs.

    Data availability

    Data available on request by contacting both the correspondent Author ([email protected]) and the New Heaven Reef Conservation Program ([email protected]).
    ReferencesRotjan, R. & Lewis, S. Impact of coral predators on tropical reefs. Mar. Ecol. Prog. Ser. 367, 73–91. https://doi.org/10.3354/meps07531 (2008).
    Google Scholar 
    Lenihan, H. S., Holbrook, S. J., Schmitt, R. J. & Brooks, A. J. Influence of corallivory, competition, and habitat structure on coral community shifts. Ecology 92(10), 1959–1971. https://doi.org/10.1890/11-0108.1 (2011).
    Google Scholar 
    Ladd, M. C. & Shantz, A. A. Trophic interactions in coral reef restoration: A review. Food Webs. 24, e00149. https://doi.org/10.1016/j.fooweb.2020.e00149 (2020).
    Google Scholar 
    Cumming, G. S., Einarsson, L. B. & Jones, G. P. Crown-of-thorns starfish promote additional fine-grained habitat fragmentation in a coral reef ecosystem. Landsc. Ecol. 40 (5), 95. https://doi.org/10.1007/s10980-025-02107-y (2025).
    Google Scholar 
    Nicolet, K. J., Hoogenboom, M. O., Gardiner, N. M., Pratchett, M. S. & Willis, B. L. The corallivorous invertebrate Drupella aids in transmission of brown band disease on the Great Barrier Reef. Coral Reefs 32(2), 585–595. https://doi.org/10.1007/s00338-013-1010-8 (2013).
    Google Scholar 
    Sussman, S. W. & Siegal, W. S. Informational influence in organizations: an integrated approach to knowledge adoption. Inf. Syst. Res. 14 (1), 47–65. https://doi.org/10.1287/isre.14.1.47.14767 (2003).
    Google Scholar 
    Turner, S. Spatial variability in the abundance of the corallivorous gastropod Drupella cornus. Coral Reefs 13, 41–48. https://doi.org/10.1007/BF00426434 (1994).
    Google Scholar 
    Zhang, F., Jia, X., Lin, Z., Jiang, Y. & Qu, M. The outbreak of drupella snails and its catastrophic effects on coral reefs: a comprehensive review. Front. Mar. Sci. 10, 1290001. https://doi.org/10.3389/fmars.2023.1290001 (2024).
    Google Scholar 
    Ayling, A. M. & Ayling, A. L. Ningaloo Marine Park: Preliminary Fish Density Assessment and Habitat Survey (Sea Research, 1987).Taylor, J. D. & Reid, D. G. The abundance and trophic classification of molluscs upon coral reefs in the Sudanese Red Sea. J. Nat. Hist. 18(2), 175–209. https://doi.org/10.1080/00222938400770151 (1984).
    Google Scholar 
    Johnson, M. S. & Cumming, R. L. Genetic distinctness of three widespread and morphologically variable species of Drupella (Gastropoda, Muricidae). Coral Reefs 14(2), 71–78. https://doi.org/10.1007/BF00303426 (1995).
    Google Scholar 
    Claremont, M., Reid, D. G. & Williams, S. T. Evolution of corallivory in the gastropod genus Drupella. Coral Reefs. 30, 977–990. https://doi.org/10.1007/s00338-011-0788-5 (2011).
    Google Scholar 
    Kita, M. et al. Feeding attractants for the muricid gastropod Drupella cornus, a coral predator. Tetrahedron Lett. 46(49), 8583–8585. https://doi.org/10.1016/j.tetlet.2005.09.182 (2005).
    Google Scholar 
    Bessey, C., Babcock, R. C., Thomson, D. P. & Haywood, M. D. E. Outbreak densities of the coral predator Drupella in relation to in situ Acropora growth rates on Ningaloo Reef, Western Australia. Coral Reefs 37(4), 985–993. https://doi.org/10.1007/s00338-018-01748-7 (2018).
    Google Scholar 
    Morton, B., Blackmore, G. & Kwok, C. T. Corallivory and prey choice by Drupella rugosa (Gastropoda:Muricidae) in Hong Kong. J. Molluscan Stud. 68(3), 217–223. https://doi.org/10.1093/mollus/68.3.217 (2002).
    Google Scholar 
    Moerland, M. S., Scott, C. M. & Hoeksema, B. W. Prey selection of corallivorous muricids at Koh Tao (Gulf of Thailand) four years after a major coral bleaching event. Contrib. Zool. 85 (3), 291–309. https://doi.org/10.1163/18759866-08503003 (2016).
    Google Scholar 
    Scott, C. M., Mehrotra, R., Cabral, M. & Arunrugstichai, S. Changes in hard coral abundance and composition on Koh Tao, Thailand, 2006–2014. Coast Ecosyst. 4, 26–38 (2017).
    Google Scholar 
    Hsieh, H. J. et al. Establishment of a no-take area (NTA) could not guarantee the preservation of coral communities in Chinwan inner Bay, Penghu, Taiwan. Zool. Stud. 50, 443–453 (2011).
    Google Scholar 
    Kaullysing, D., Taleb-Hossenkhan, N., Kulkarni, B. G. & Bhagooli, R. A first field report of various coral-eating gastropods and associated infestations around Mauritius Island, Western Indian ocean. West Indian Ocean. J. Mar. Sci. (1), 73–75 (2017).Bruckner, A. W. Priorities for Effective Management of Coral Diseases (U.S. Department of Commerce, National Oceanic and Atmospheric Administration, National Marine Fisheries Service, 2002).Samsuri, A. N. et al. The effectiveness of Trapezia cymodoce in defending its host coral Pocillopora acuta against corallivorous Drupella. Mar. Biol. 165(4), 70. https://doi.org/10.1007/s00227-018-3330-2 (2018).
    Google Scholar 
    Forde, F. J. M. Populations, behaviour and effects of drupella cornus on the Ningaloo reef. Conserv. Land. Manage. 92, 45–50 (1992).
    Google Scholar 
    Schoepf, V., Herler, J. & Zuschin, M. Microhabitat use and prey selection of the coral-feeding snail Drupella cornus in the northern Red Sea. Hydrobiologia 641(1), 45–57. https://doi.org/10.1007/s10750-009-0053-x (2010).
    Google Scholar 
    Hoeksema, B. W., Scott, C. & True, J. D. Dietary shift in corallivorous drupella snails following a major bleaching event at Koh Tao, Gulf of Thailand. Coral Reefs. 32 (2), 423–428. https://doi.org/10.1007/s00338-012-1005-x (2013).
    Google Scholar 
    Lei, X. et al. Spatial variability In the abundance and prey selection of the corallivorous snail drupella spp. In the southeastern Hainan Island, China. Front. Mar. Sci. 9, 990113. https://doi.org/10.3389/fmars.2022.990113 (2022).
    Google Scholar 
    Moore, R. J. Is Acanthaster planci an r-strategist?. Nature 271(5640), 56–57. https://doi.org/10.1038/271056a0 (1978).
    Google Scholar 
    Levitan, D. R. ‘The ecology of fertilization in free-spawning invertebrates.’ In Ecology of Marine Invertebrate Larvae 123–156 (CRC, 2020).
    Google Scholar 
    Pechenik, J. On the advantages and disadvantages of larval stages in benthic marine invertebrate life cycles. Mar. Ecol. Prog Ser. 177, 269–297. https://doi.org/10.3354/meps177269 (1999).
    Google Scholar 
    Underwood, A. J. Supply-side ecology: The natural and consequences of variations in recruitment of intertidal organisms. Mar. Community Ecol. (2001).Kitamura, T., Shigematsu, Y., Iwai, T., Miura, C. & Miura, T. The spawning season of Drupella fragum in southwestern Shikoku. Biogeography. https://doi.org/10.11358/biogeo.24.32 (2022).
    Google Scholar 
    Scott, C. M., Mehrotra, R., Hein, M. Y., Moerland, M. S. & Hoeksema, B. W. Population dynamics of corallivores (Drupella and Acanthaster) on coral reefs of Koh Tao, a diving destination in the Gulf of Thailand. Raffles Bull. Zool. 65 (2017).Canteri, B. Investigating climate change and nutrient pollution effects on Drupella rugosa coral reef destruction in Koh-Tao, Thailand. Plymouth Stud. Sci. 17 (2), 14. https://doi.org/10.70156/1754-2383.1494 (2024).
    Google Scholar 
    Pratchett, M. S. & Cumming, G. S. Managing cross-scale dynamics in marine conservation: Pest irruptions and lessons from culling of crown-of-thorns starfish (Acanthaster spp). Biol. Conserv. 238, 108211. https://doi.org/10.1016/j.biocon.2019.108211 (2019).
    Google Scholar 
    Giakoumi, S. et al. Management priorities for marine invasive species. Sci. Total Environ. 688, 976–982. https://doi.org/10.1016/j.scitotenv.2019.06.282 (2019).
    Google Scholar 
    Simberloff, D. Maintenance management and eradication of established aquatic invaders. Hydrobiologia 848(9), 2399–2420. https://doi.org/10.1007/s10750-020-04352-5 (2021).
    Google Scholar 
    Thresher, R. E. & Kuris, A. M. Options for managing invasive marine species. Biol. Invasions 6(3), 295–300. https://doi.org/10.1023/B:BINV.0000034598.28718.2e (2004).
    Google Scholar 
    Baruffaldi, M. et al. Coral health status before and after the tourism halt caused by the COVID-19 pandemic in Koh Tao (Thailand). Coral Reefs https://doi.org/10.1007/s00338-025-02706-w (2025).
    Google Scholar 
    Saponari, L., Dehnert, I., Galli, P. & Montano, S. Assessing population collapse of Drupella spp. (Mollusca: Gastropoda) 2 years after a coral bleaching event in the Republic of Maldives. Hydrobiologia 848(11), 2653–2666. https://doi.org/10.1007/s10750-021-04546-5 (2021).
    Google Scholar 
    Wickham, H. ggplot2: Elegant Graphics for Data Analysis. (Springer-Verlag New York, 2016). https://ggplot2.tidyverse.orgFox, J. & Weisberg, S. An R Companion to Applied Regression, Third (Sage, 2019). https://www.john-fox.ca/CompanionKassambara, A. ‘rstatix: pipe-friendly framework for basic statistical tests (R package version 0.7.0)’. (2021). https://CRAN.R-project.org/package=rstatixCore Team, R. R: a language and environment for statistical com- puting. R Foundation for Statistical Computing, Vienna, Austria. (2025). https://www.R-project.orgFontoura-da-Silva, V., Cardoso, R. S. & Caetano, C. H. S. Mark–recapture versus length–frequency based methods: evaluation using a marine gastropod as a model. J. Exp. Mar. Biol. Ecol. 474, 171–179. https://doi.org/10.1016/j.jembe.2015.10.013 (2016).
    Google Scholar 
    Rempel, H. S., Bodwin, K. N. & Ruttenberg, B. I. Impacts of parrotfish predation on a major reef-building coral: quantifying healing rates and thresholds of coral recovery. Coral Reefs 39(5), 1441–1452. https://doi.org/10.1007/s00338-020-01977-9 (2020).
    Google Scholar 
    Cumming, R. L. Case Study: Impact of Drupella spp. On reef-building Corals of the Great Barrier Reef (Great Barrier Reef Marine Park Authority, 2009).Schneider, C. A., Rasband, W. S. & Eliceiri, K. W. NIH Image to ImageJ: 25 years of image analysis. Nat. Methods 9(7), 671–675. https://doi.org/10.1038/nmeth.2089 (2012).
    Google Scholar 
    Tiddy, I. C. et al. Outplanting of branching acropora enhances recolonization of a fish species and protects massive corals from predation. Coral Reefs. 40 (5), 1549–1561. https://doi.org/10.1007/s00338-021-02147-1 (2021).
    Google Scholar 
    Schopmeyer, S. A. & Lirman, D. Occupation dynamics and impacts of damselfish territoriality on recovering populations of the threatened Staghorn coral, Acropora cervicornis. PLoS ONE. 10 (11), e0141302. https://doi.org/10.1371/journal.pone.0141302 (2015).
    Google Scholar 
    Losey, G. S. Jr. The ecological importance of cleaning symbiosis. Copeia https://doi.org/10.2307/1442741 (1972).
    Google Scholar 
    Sam, S. et al. First observation of Drupella rugosa egg capsules on scleractinian coral Pocillopora damicornis. Bull. Mar. Sci. 92(3), 353–354. https://doi.org/10.5343/bms.2016.1062 (2016).
    Google Scholar 
    Sam, S. Q. et al. Egg capsules and veligers of the corallivorous muricid gastropod Drupella rugosa (Born, 1778). Invertebr Reprod. Dev. 61(3), 164–171. https://doi.org/10.1080/07924259.2017.1315343 (2017).
    Google Scholar 
    Ben-Ari, H., Paz, M. & Sher, D. The chemical armament of reef-building corals: inter- and intra-specific variation and the identification of an unusual actinoporin in Stylophora pistilata. Sci. Rep. 8 (1), 1–13. https://doi.org/10.1038/s41598-017-18355-1 (2018).
    Google Scholar 
    Potts, D. C. Suppression of coral populations by filamentous algae within damselfish territories. J. Exp. Mar. Biol. Ecol. 28(3), 207–216. https://doi.org/10.1016/0022-0981(77)90092-2 (1977).
    Google Scholar 
    Reinthal, P. N. & Macintyre, I. G. Spatial and Temporal variations in grazing pressure by herbivorous fishes: tobacco Reef, Belize. Atoll Res. Bull. 425, 1–11. https://doi.org/10.5479/si.00775630.425.1 (1994).
    Google Scholar 
    Monchanin, C., Desmolles, M. & Mehrotra, R. Homogenization and distinction of coral recruit communities between natural and artificial substrates at Koh Tao a decade after deployment. Aquat. Ecol. 59 (2), 597–608. https://doi.org/10.1007/s10452-025-10182-1 (2025).
    Google Scholar 
    Zavaleta, E. S., Hobbs, R. J. & Mooney, H. A. Viewing invasive species removal in a whole-ecosystem context. Trends Ecol. Evol. 16, 454–459. https://doi.org/10.1016/S0169-5347(01)02194-2 (2001).
    Google Scholar 
    Côté, I. M., Akins, L., Underwood, E., Curtis-Quick, J. & Green, S. J. Setting the record straight on invasive lionfish control: culling works. PeerJ Prepr. 2, e398v1 (2014).
    Google Scholar 
    Weterings, R. ‘A GIS-based assessment of threats to the natural environment on Koh Tao, Thailand. Agric. Nat. Resour. 45(4), 743–755 (2011).
    Google Scholar 
    Haslam, V. M., Bessey, C., Chaplin, J. A. & van Keulen, M. Evidence of corallivorous gastropod drupella cornus breeding on the higher latitude reefs of Rottnest Island (32° S), Western Australia. Mar. Biol. 171 (1), 28. https://doi.org/10.1007/s00227-023-04352-8 (2024).
    Google Scholar 
    Costello, M. J. et al. Biological and ecological traits of marine species. PeerJ 3, e1201. https://doi.org/10.7717/peerj.1201 (2015).
    Google Scholar 
    Marchesi, V. et al. A baseline for the conservation of the native and protected Centrostephanus longispinus (Philippi, 1845) and the management of the invasive Diadema setosum (Leske, 1778) (Echinoidea: diadematidae) in the mediterranean sea. Aquat. Conserv. Mar. Freshw. Ecosyst. 35 (5), 1–12. https://doi.org/10.1002/aqc.70155 (2025).
    Google Scholar 
    Green, S. J. & Grosholz, E. D. Functional eradication as a framework for invasive species control. Front. Ecol. Environ. 19 (2), 98–107. https://doi.org/10.1002/fee.2277 (2021).
    Google Scholar 
    Hulme, P. E. Beyond control: wider implications for the management of biological invasions. J. Appl. Ecol. 43 (5), 835–847. https://doi.org/10.1111/j.1365-2664.2006.01227.x (2006).
    Google Scholar 
    Pluess, T. et al. When are eradication campaigns successful? A test of common assumptions. Biol. Invasions 14(7), 1365–1378. https://doi.org/10.1007/s10530-011-0160-2 (2012).
    Google Scholar 
    Osborne, S. & Williams, M. R. A preliminary summary of the effects of hand removal of Drupella cornus on Ningaloo Reef. In Drupella cornus: A Synopsis, 83–90 (1992).Williams, D. E., Miller, M. W., Bright, A. J. & Cameron, C. M. Removal of corallivorous snails as a proactive tool for the conservation of acroporid corals. PeerJ 2, e680. https://doi.org/10.7717/peerj.680 (2014).
    Google Scholar 
    Williams, S. L. & Grosholz, E. D. The invasive species challenge in estuarine and coastal environments: marrying management and science. Estuaries Coasts 31(1), 3–20. https://doi.org/10.1007/s12237-007-9031-6 (2008).
    Google Scholar 
    Ojaveer, H. et al. Classification of non-indigenous species based on their impacts: considerations for application in marine management. PLoS Biol. 13 (4), e1002130. https://doi.org/10.1371/journal.pbio.1002130 (2015).
    Google Scholar 
    Yamaguchi, M. Acanthaster planci infestations of reefs and coral assemblages in japan: a retrospective analysis of control efforts. Coral Reefs. 5, 23–30. https://doi.org/10.1007/BF00302168 (1986).
    Google Scholar 
    Rivera-Posada, J. Size-related variation in arm damage frequency in the crown-of-thorns sea star, Acanthaster planci. J. Coast Life Med. https://doi.org/10.12980/JCLM.2.2014J52 (2014).
    Google Scholar 
    Westcott, D. A. et al. Relative efficacy of three approaches to mitigate Crown-of-Thorns starfish outbreaks on australia’s great barrier reef. Sci. Rep. 10 (1), 12594. https://doi.org/10.1038/s41598-020-69466-1 (2020).
    Google Scholar 
    Strand, H. K., Christie, H., Fagerli, C. W., Mengede, M. & Moy, F. Optimizing the use of quicklime (CaO) for sea urchin management—a lab and field study. Ecol. Eng. 143, 100018. https://doi.org/10.1016/j.ecoena.2020.100018 (2020).
    Google Scholar 
    Christie, H. et al. Successful large-scale and long-term Kelp forest restoration by culling sea urchins with quicklime and supported by crab predation. Mar. Biol. 171 (11), 211. https://doi.org/10.1007/s00227-024-04540-0 (2024).
    Google Scholar 
    Miller, M. W. Corallivorous snail removal: evaluation of impact on Acropora palmata. Coral Reefs 19(3), 293–295. https://doi.org/10.1007/PL00006963 (2001).
    Google Scholar 
    Tracey, S. R. et al. Systematic culling controls a climate driven, habitat modifying invader. Biol. Invasions. 17 (6), 1885–1896. https://doi.org/10.1007/s10530-015-0845-z (2015).
    Google Scholar 
    Ling, S. D. & Keane, J. P. Resurvey of the longspined sea urchin (Centrostephanus rodgersii) and associated barren reef in Tasmania. (2018). https://doi.org/10.13140/RG.2.2.16363.80162Sanderson, J. C., Ling, S. D., Dominguez, J. G. & Johnson, C. R. Limited effectiveness of divers to mitigate barrens formation by culling sea urchins while fishing for abalone. Mar. Freshw. Res. 67(1), 84–95. https://doi.org/10.1071/MF14255 (2015).
    Google Scholar 
    Usseglio, P., Selwyn, J. D., Downey-Wall, A. M. & Hogan, J. D. Effectiveness of removals of the invasive lionfish: how many dives are needed to deplete a reef? PeerJ 5, e3043. https://doi.org/10.7717/peerj.3043 (2017).
    Google Scholar 
    Morris, J. A., Sullivan, C. V. & Govoni, J. J. Oogenesis and spawn formation in the invasive lionfish, Pterois miles and Pterois volitans. Sci. Mar. 75(1), 147–154. https://doi.org/10.3989/scimar.2011.75n1147 (2011).
    Google Scholar 
    Bohn, K., Richardson, C. A. & Jenkins, S. R. The importance of larval supply, larval habitat selection and post-settlement mortality in determining intertidal adult abundance of the invasive gastropod Crepidula fornicata. J. Exp. Mar. Biol. Ecol. 440, 132–140. https://doi.org/10.1016/j.jembe.2012.12.008 (2013).
    Google Scholar 
    Phillips, W. N. Tourism threats to coral reef resilience at Koh Sak, Pattaya Bay. Environ. Nat. Resour. J. https://doi.org/10.14456/ENNRJ.2015.3 (2015).
    Google Scholar 
    Artificial intelligence in invasive species management: Transforming detection and response. Trends Anim. Plant. Sci. 4, 82–96. https://doi.org/10.62324/TAPS/2024.050 (2024).Katsanevakis, S. et al. Marine invasive alien species in europe: 9 years after the IAS regulation. Front. Mar. Sci. 10, 1271755. https://doi.org/10.3389/fmars.2023.1271755 (2023).
    Google Scholar 
    Holmes, R. B., Matchette, S. R. & Herbert-Read, J. E. Citizen science reveals relationships between human hunting pressure and the abundance and behaviour of invasive lionfish (Pterois spp.). Biol. Invasions 27(6), 141. https://doi.org/10.1007/s10530-025-03596-3 (2025).
    Google Scholar 
    Malpica-Cruz, L. et al. Trying to collapse a population for conservation: commercial trade of a marine invasive species by artisanal fishers. Rev. Fish. Biol. Fish. 31 (3), 667–683. https://doi.org/10.1007/s11160-021-09660-0 (2021).
    Google Scholar 
    Download referencesAcknowledgementsWe would like to thank New Heaven Reef Conservation and above all Kirsty Magson, the program manager, for providing part of the data and making this study possible. We are also deeply grateful to all the volunteers who, over the years, have contributed to the collection of data and to the implementation of the research.FundingNo Funding.Author informationAuthors and AffiliationsNew Heaven Reef Conservation Program, 48 Moo 3, Chalok Ban Kao, Koh Tao, 84360, ThailandBaruffaldi MatildeDepartment of Life and Environmental Sciences (DiSVA), Università Politecnica delle Marche, Via Brecce Bianche s.n.c, 60131, Ancona, ItalyBaruffaldi Matilde, Roveta Camilla, Tonolini Rosita, Pulido Mantas Torcuato & Cristina Gioia Di CamilloNational Biodiversity Future Center (NBFC), Piazza Marina 61, 90133, Palermo, ItalyRoveta Camilla, Pulido Mantas Torcuato & Cristina Gioia Di CamilloConsorzio Nazionale Interuniversitario per le Scienze del Mare (CoNISMa), Piazzale Flaminio 9, 00196, Rome, ItalyCristina Gioia Di CamilloAuthorsBaruffaldi MatildeView author publicationsSearch author on:PubMed Google ScholarRoveta CamillaView author publicationsSearch author on:PubMed Google ScholarTonolini RositaView author publicationsSearch author on:PubMed Google ScholarPulido Mantas TorcuatoView author publicationsSearch author on:PubMed Google ScholarCristina Gioia Di CamilloView author publicationsSearch author on:PubMed Google ScholarContributionsDi Camillo CG and Baruffaldi M contributed to the study conception and design. Baruffaldi M and Di Camillo CG wrote the first draft of the paper and provided figures. Baruffaldi M and Tonolini R performed samplings and collected data. Roveta C, Baruffaldi M, Pulido Mantas T analyzed data. All authors contributed to improve and revise the manuscript.Corresponding authorCorrespondence to
    Cristina Gioia Di Camillo.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-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 articleMatilde, B., Camilla, R., Rosita, T. et al. Short- and long-term effects of culling invasive corallivorous gastropods.
    Sci Rep (2025). https://doi.org/10.1038/s41598-025-32915-wDownload citationReceived: 15 September 2025Accepted: 13 December 2025Published: 19 December 2025DOI: https://doi.org/10.1038/s41598-025-32915-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
    KeywordsBiological invasionsRemovalMuricidaeControl of pestsSurge More