More stories

  • in

    Ecological insights from three decades of forest biodiversity experiments

    AbstractForest biodiversity experiments test how species diversity affects forest ecosystem functioning, typically in terms of forest productivity. In this Review, we discuss key findings from these experiments and put them into context with observational studies from forests. Experimental studies can reveal causal effects of biodiversity on ecosystem functioning, which is extremely challenging in observational studies. The past three decades of experimental research show that increasing tree diversity can promote a multitude of ecosystem functions through resource partitioning, abiotic and biotic facilitation, and other species interactions. The longest-running experiments show that these relationships strengthen over time, and comparative work in natural or planted forests suggests that these effects are likely to persist. Moreover, diversity at other trophic levels can strongly mediate tree diversity effects on forest productivity. New experiments that manipulate both tree diversity and the diversity of other trophic levels as orthogonal treatments are needed to investigate causality in these interactions. Furthermore, experiments crossing tree diversity with global change factors are necessary to understand the context-dependency of tree diversity–ecosystem functioning relationships under global change. Finally, combining insights from observational studies and experiments can help biodiversity–ecosystem function research to inform restoration and forest management targets of the Global Biodiversity Framework.

    Access through your institution

    Buy or subscribe

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

    Access options

    Access through your institution

    Subscribe to this journal

    Receive 12 digital issues and online access to articles

    $119.00 per year
    only $9.92 per issue

    Learn more

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

    Additional access options:

    Log in

    Learn about institutional subscriptions

    Read our FAQs

    Contact customer support

    Fig. 1: The locations and types of 45 forest biodiversity experiments covering 72 sites globally.Fig. 2: Potential mechanisms underlying the tree diversity effect over time.Fig. 3: Trophic mediation of tree diversity effects via interactions with and among higher trophic levels.Fig. 4: Experimental and observational BEF approaches.

    ReferencesCardinale, B. J. et al. Biodiversity loss and its impact on humanity. Nature 486, 59–67 (2012).Article 
    CAS 

    Google Scholar 
    IPBES. Global Assessment Report on Biodiversity and Ecosystem Services of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services (IPBES Secretariat, 2019).Loreau, M. et al. Ecology — biodiversity and ecosystem functioning: current knowledge and future challenges. Science 294, 804–808 (2001).Article 
    CAS 

    Google Scholar 
    Balvanera, P. et al. Quantifying the evidence for biodiversity effects on ecosystem functioning and services. Ecol. Lett. 9, 1146–1156 (2006).Article 

    Google Scholar 
    Liu, X. et al. Tree species richness increases ecosystem carbon storage in subtropical forests. Proc. Biol. Sci. 285, 20181240 (2018).
    Google Scholar 
    Liu, Y. P. et al. Biodiversity and productivity in eastern US forests. Proc. Natl Acad. Sci. USA 121, e2314231121 (2024).Article 
    CAS 

    Google Scholar 
    Isbell, F. et al. High plant diversity is needed to maintain ecosystem services. Nature 477, 199–202 (2011).Article 
    CAS 

    Google Scholar 
    Hooper, D. U. et al. Effects of biodiversity on ecosystem functioning: a consensus of current knowledge. Ecol. Monogr. 75, 3–35 (2005).Article 

    Google Scholar 
    Chen, C., Xiao, W. & Chen, H. Y. H. Meta-analysis reveals global variations in plant diversity effects on productivity. Nature 638, 435–440 (2025).Article 
    CAS 

    Google Scholar 
    Hector, A. et al. Plant diversity and productivity experiments in European grasslands. Science 286, 1123–1127 (1999).Article 
    CAS 

    Google Scholar 
    Huston, M. A. et al. No consistent effect of plant diversity on productivity. Science 289, 1255–1255 (2000).Article 
    CAS 

    Google Scholar 
    Grace, J. B. et al. Integrative modelling reveals mechanisms linking productivity and plant species richness. Nature 529, 390–393 (2016).Article 
    CAS 

    Google Scholar 
    Duffy, J. E., Godwin, C. M. & Cardinale, B. J. Biodiversity effects in the wild are common and as strong as key drivers of productivity. Nature 549, 261–264 (2017).Article 
    CAS 

    Google Scholar 
    Tilman, D., Wedin, D. & Knops, J. Productivity and sustainability influenced by biodiversity in grassland ecosystems. Nature 379, 718–720 (1996).Article 
    CAS 

    Google Scholar 
    Tilman, D. et al. Diversity and productivity in a long-term grassland experiment. Science 294, 843–845 (2001).Article 
    CAS 

    Google Scholar 
    Scherer-Lorenzen, M. et al. in Forest Diversity and Function: Temperate and Boreal Systems (eds Scherer-Lorenzen, M., Körner, C. & Schulze, E.-D.) 347–376 (Springer, 2005).Paquette, A. et al. A million and more trees for science. Nat. Ecol. Evol. 2, 763–766 (2018).Article 

    Google Scholar 
    Koricheva, J. et al. Using long-term tree diversity experiments to explore the mechanisms of temporal shifts in forest ecosystem functioning. Oikos 2025, e10872 (2025).Article 
    CAS 

    Google Scholar 
    Liang, J. J. et al. Positive biodiversity-productivity relationship predominant in global forests. Science 354, aaf8957 (2016).Article 

    Google Scholar 
    Feng, Y. et al. Multispecies forest plantations outyield monocultures across a broad range of conditions. Science 376, 865–868 (2022).Article 
    CAS 

    Google Scholar 
    Paquette, A. & Messier, C. The effect of biodiversity on tree productivity: from temperate to boreal forests. Glob. Ecol. Biogeogr. 20, 170–180 (2011).Article 

    Google Scholar 
    Zhang, Y., Chen, H. Y. H. & Reich, P. B. Forest productivity increases with evenness, species richness and trait variation: a global meta-analysis. J. Ecol. 100, 742–749 (2012).Article 

    Google Scholar 
    Brown, A. H. F. in The Ecology of Mixed-Species Stands of Trees (eds Cannell, M. G. R. et al.) 125–150 (Blackwell, 1992).Dee, L. E. et al. Clarifying the effect of biodiversity on productivity in natural ecosystems with longitudinal data and methods for causal inference. Nat. Commun. 14, 2607 (2023).Article 
    CAS 

    Google Scholar 
    Jewell, M. D. et al. Partitioning the effect of composition and diversity of tree communities on leaf litter decomposition and soil respiration. Oikos 126, 959–971 (2017).Article 
    CAS 

    Google Scholar 
    Jewell, M. D., Shipley, B., Paquette, A., Messier, C. & Reich, P. B. A traits-based test of the home-field advantage in mixed-species tree litter decomposition. Ann. Botany 116, 781–788 (2015).Article 

    Google Scholar 
    Berthelot, S. et al. Tree diversity reduces the risk of bark beetle infestation for preferred conifer species, but increases the risk for less preferred hosts. J. Ecol. 109, 2649–2661 (2021).Article 

    Google Scholar 
    Schnabel, F. et al. Species richness stabilizes productivity via asynchrony and drought-tolerance diversity in a large-scale tree biodiversity experiment. Sci. Adv. 7, eabk1643 (2021).Article 

    Google Scholar 
    Blondeel, H. et al. Tree diversity reduces variability in sapling survival under drought. J. Ecol. 112, 1164–1180 (2024).Article 

    Google Scholar 
    Williams, L. J. et al. Remote spectral detection of biodiversity effects on forest biomass. Nat. Ecol. Evol. 5, 46–54 (2021).Article 

    Google Scholar 
    Ray, T. et al. Tree diversity increases productivity through enhancing structural complexity across mycorrhizal types. Sci. Adv. 9, eadi2362 (2023).Article 

    Google Scholar 
    Williams, L. J., Paquette, A., Cavender-Bares, J., Messier, C. & Reich, P. B. Spatial complementarity in tree crowns explains overyielding in species mixtures. Nat. Ecol. Evol. 1, 0063 (2017).Article 

    Google Scholar 
    Guzmán, Q. J. A., Park, M. H., Williams, L. J. & Cavender-Bares, J. Seasonal structural stability promoted by forest diversity and composition explains overyielding. Ecology 106, e70055 (2025).Article 

    Google Scholar 
    Deng, X. et al. Forest biodiversity increases productivity via complementarity from greater canopy structural complexity. Proc. Natl Acad. Sci. USA 122, e2506750122 (2025).Article 
    CAS 

    Google Scholar 
    Williams, L. J. et al. Enhanced light interception and light use efficiency explain overyielding in young tree communities. Ecol. Lett. 24, 996–1006 (2021).Article 

    Google Scholar 
    Sapijanskas, J., Paquette, A., Potvin, C., Kunert, N. & Loreau, M. Tropical tree diversity enhances light capture through crown plasticity and spatial and temporal niche differences. Ecology 95, 2479–2492 (2014).Article 

    Google Scholar 
    Grossman, J. J., Cavender-Bares, J. & Hobbie, S. E. Functional diversity of leaf litter mixtures slows decomposition of labile but not recalcitrant carbon over two years. Ecol. Monogr. 90, e01407 (2020).Article 

    Google Scholar 
    Bryant, R. L. et al. Independent effects of tree diversity on aboveground and soil carbon pools after six years of experimental afforestation. Ecol. Appl. 34, e3042 (2024).Article 

    Google Scholar 
    Archambault, C. et al. Evergreenness influences fine root growth more than tree diversity in a common garden experiment. Oecologia 189, 1027–1039 (2019).Article 

    Google Scholar 
    Sun, Z. et al. Positive effects of tree species richness on fine-root production in a subtropical forest in SE-China. J. Plant Ecol. 10, 146–157 (2017).Article 

    Google Scholar 
    Li, Y. et al. Multitrophic arthropod diversity mediates tree diversity effects on primary productivity. Nat. Ecol. Evol. 7, 832–840 (2023).Article 

    Google Scholar 
    Li, Y. et al. Plant diversity enhances ecosystem multifunctionality via multitrophic diversity. Nat. Ecol. Evol. 8, 2037–2047 (2024).Article 

    Google Scholar 
    Singavarapu, B. et al. Influence of tree mycorrhizal type, tree species identity, and diversity on forest root-associated mycobiomes. New Phytol. 242, 1691–1703 (2024).Article 
    CAS 

    Google Scholar 
    Laforest-Lapointe, I., Paquette, A., Messier, C. & Kembel, S. W. Leaf bacterial diversity mediates plant diversity and ecosystem function relationships. Nature 546, 145–147 (2017).Article 
    CAS 

    Google Scholar 
    Depauw, L. et al. Enhancing tree performance through species mixing: review of a quarter-century of treedivnet experiments reveals research gaps and practical insights. Curr. Forestry Rep. 10, 1–20 (2024).Article 

    Google Scholar 
    Urgoiti, J. et al. No complementarity no gain — net diversity effects on tree productivity occur once complementarity emerges during early stand development. Ecol. Lett. 25, 851–862 (2022).Article 

    Google Scholar 
    Grossman, J. J., Cavender-Bares, J., Hobbie, S. E., Reich, P. B. & Montgomery, R. A. Species richness and traits predict overyielding in stem growth in an early-successional tree diversity experiment. Ecology 98, 2601–2614 (2017).Article 

    Google Scholar 
    Tobner, C. M. et al. Functional identity is the main driver of diversity effects in young tree communities. Ecol. Lett. 19, 638–647 (2016).Article 

    Google Scholar 
    Chen, Y. et al. Directed species loss reduces community productivity in a subtropical forest biodiversity experiment. Nat. Ecol. Evol. 4, 550–559 (2020).Article 

    Google Scholar 
    Fichtner, A. et al. Neighbourhood interactions drive overyielding in mixed-species tree communities. Nat. Commun. 9, 1144 (2018).Article 

    Google Scholar 
    Huang, Y. et al. Impacts of species richness on productivity in a large-scale subtropical forest experiment. Science 362, 80–83 (2018).Article 
    CAS 

    Google Scholar 
    Schnabel, F. et al. Drivers of productivity and its temporal stability in a tropical tree diversity experiment. Glob. Change Biol. 25, 4257–4272 (2019).Article 

    Google Scholar 
    Veryard, R. et al. Positive effects of tree diversity on tropical forest restoration in a field-scale experiment. Sci. Adv. 9, eadf0938 (2023).Article 

    Google Scholar 
    Xu, Q. et al. Consistently positive effect of species diversity on ecosystem, but not population, temporal stability. Ecol. Lett. 24, 2256–2266 (2021).Article 

    Google Scholar 
    Loreau, M. & Hector, A. Partitioning selection and complementarity in biodiversity experiments. Nature 412, 72–76 (2001).Article 
    CAS 

    Google Scholar 
    Haggar, J. P. & Ewel, J. J. Primary productivity and resource partitioning in model tropical ecosystems. Ecology 78, 1211–1221 (1997).Article 

    Google Scholar 
    Chen, C. et al. Understory shrub diversity: equally vital as overstory tree diversity to promote forest productivity. Natl Sci. Rev. 12, nwaf093 (2025).Article 

    Google Scholar 
    Bongers, F. J. et al. Functional diversity effects on productivity increase with age in a forest biodiversity experiment. Nat. Ecol. Evol. 5, 1594–1603 (2021).Article 

    Google Scholar 
    Ferlian, O. et al. Mycorrhiza in tree diversity–ecosystem function relationships: conceptual framework and experimental implementation. Ecosphere 9, e02226 (2018).Article 

    Google Scholar 
    Tobner, C. M., Paquette, A., Reich, P. B., Gravel, D. & Messier, C. Advancing biodiversity–ecosystem functioning science using high-density tree-based experiments over functional diversity gradients. Oecologia 174, 609–621 (2014).Article 

    Google Scholar 
    Cavender-Bares, J. et al. Forest and Biodiversity 2: a tree diversity experiment to understand the consequences of multiple dimensions of diversity and composition for long-term ecosystem function and resilience. Methods Ecol. Evol. 15, 2400–2414 (2024).Article 

    Google Scholar 
    Verheyen, K. et al. Assessment of the functional role of tree diversity: the multi-site FORBIO experiment. Plant Ecol. Evol. 146, 26–35 (2013).Article 

    Google Scholar 
    Bruelheide, H. et al. Designing forest biodiversity experiments: general considerations illustrated by a new large experiment in subtropical China. Methods Ecol. Evol. 5, 74–89 (2014).Article 

    Google Scholar 
    Tang, T. et al. Identifying seed families with high mixture performance in a subtropical forest biodiversity experiment. New Phytol. 246, 2537–2550 (2025).Article 

    Google Scholar 
    Chen, C., Bongers, F. J., Schmid, B., Ma, K. & Liu, X. Ecosystem consequences of functional diversity in forests and implications for restoration. New Phytol. 247, 1081–1097 (2025).Article 

    Google Scholar 
    Hobbie, S. E. et al. Tree species effects on decomposition and forest floor dynamics in a common garden. Ecology 87, 2288–2297 (2006).Article 

    Google Scholar 
    Cavender-Bares, J., Kozak, K. H., Fine, P. V. A. & Kembel, S. W. The merging of community ecology and phylogenetic biology. Ecol. Lett. 12, 693–715 (2009).Article 

    Google Scholar 
    Grossman, J. J., Cavender-Bares, J., Reich, P. B., Montgomery, R. A. & Hobbie, S. E. Neighborhood diversity simultaneously increased and decreased susceptibility to contrasting herbivores in an early stage forest diversity experiment. J. Ecol. 107, 1492–1505 (2019).Article 

    Google Scholar 
    Darwin, C. On the Origin of Species by Means of Natural Selection (John Murray, 1859).Janzen, D. H. Herbivores and the number of tree species in tropical forests. Am. Naturalist 104, 501–528 (1970).Article 

    Google Scholar 
    Larkin, D. J. et al. Evolutionary history shapes grassland productivity through opposing effects on complementarity and selection. Ecology 104, e4129 (2023).Article 

    Google Scholar 
    Venail, P. et al. Species richness, but not phylogenetic diversity, influences community biomass production and temporal stability in a re-examination of 16 grassland biodiversity studies. Funct. Ecol. 29, 615–626 (2015).Article 

    Google Scholar 
    Helmus, M. R., Bland, T. J., Williams, C. K. & Ives, A. R. Phylogenetic measures of biodiversity. Am. Naturalist 169, E68–E83 (2007).Article 

    Google Scholar 
    E-Vojtkó, A., Bello, F., Lososová, Z. & Götzenberger, L. Phylogenetic diversity is a weak proxy for functional diversity but they are complementary in explaining community assembly patterns in temperate vegetation. J. Ecol. 111, 2218–2230 (2023).Article 

    Google Scholar 
    Flynn, D. F. B., Mirotchnick, N., Jain, M., Palmer, M. I. & Naeem, S. Functional and phylogenetic diversity as predictors of biodiversity–ecosystem-function relationships. Ecology 92, 1573–1581 (2011).Article 

    Google Scholar 
    Devictor, V. et al. Spatial mismatch and congruence between taxonomic, phylogenetic and functional diversity: the need for integrative conservation strategies in a changing world. Ecol. Lett. 13, 1030–1040 (2010).Article 

    Google Scholar 
    Mazel, F. et al. Prioritizing phylogenetic diversity captures functional diversity unreliably. Nat. Commun. 9, 2888 (2018).Article 

    Google Scholar 
    Ackerly, D. D. & Reich, P. B. Convergence and correlations among leaf size and function in seed plants: a comparative test using independent contrasts. Am. J. Botany 86, 1272–1281 (1999).Article 
    CAS 

    Google Scholar 
    Crisp, M. D. & Cook, L. G. Phylogenetic niche conservatism: what are the underlying evolutionary and ecological causes? New Phytol. 196, 681–694 (2012).Article 

    Google Scholar 
    Paquette, A., Joly, S. & Messier, C. Explaining forest productivity using tree functional traits and phylogenetic information: two sides of the same coin over evolutionary scale? Ecol. Evol. 5, 1774–1783 (2015).Article 

    Google Scholar 
    Le Bagousse-Pinguet, Y. et al. Phylogenetic, functional, and taxonomic richness have both positive and negative effects on ecosystem multifunctionality. Proc. Natl Acad. Sci. USA 116, 8419–8424 (2019).Article 

    Google Scholar 
    Díaz, S. et al. Functional traits, the phylogeny of function, and ecosystem service vulnerability. Ecol. Evol. 3, 2958–2975 (2013).Article 

    Google Scholar 
    Zhang, L. et al. Strong nestedness and turnover effects on stand productivity in a long-term forest biodiversity experiment. New Phytol. 245, 130–140 (2025).Article 
    CAS 

    Google Scholar 
    Boshier, D. et al. Is local best? Examining the evidence for local adaptation in trees and its scale. Environ. Evid. 4, 20 (2015).Article 

    Google Scholar 
    Luo, S. et al. Community-wide trait means and variations affect biomass in a biodiversity experiment with tree seedlings. Oikos 129, 799–810 (2020).Article 

    Google Scholar 
    Schweiger, A. K. et al. Plant spectral diversity integrates functional and phylogenetic components of biodiversity and predicts ecosystem function. Nat. Ecol. Evol. 2, 976–982 (2018).Article 

    Google Scholar 
    Felipe-Lucia, M. R. et al. Multiple forest attributes underpin the supply of multiple ecosystem services. Nat. Commun. 9, 4839 (2018).Article 

    Google Scholar 
    Kothari, S., Montgomery, R. A. & Cavender-Bares, J. Physiological responses to light explain competition and facilitation in a tree diversity experiment. J. Ecol. 109, 2000–2018 (2021).Article 

    Google Scholar 
    Zapater, M. et al. Evidence of hydraulic lift in a young beech and oak mixed forest using 18O soil water labelling. Trees 25, 885–894 (2011).Article 

    Google Scholar 
    Schnabel, F. et al. Tree diversity increases forest temperature buffering via enhancing canopy density and structural diversity. Ecol. Lett. 28, e70096 (2025).Article 

    Google Scholar 
    Salmon, Y. et al. Surrounding species diversity improves subtropical seedlings’ carbon dynamics. Ecol. Evol. 8, 7055–7067 (2018).Article 

    Google Scholar 
    Zhang, S., Landuyt, D., Verheyen, K. & De Frenne, P. Tree species mixing can amplify microclimate offsets in young forest plantations. J. Appl. Ecol. 59, 1428–1439 (2022).Article 

    Google Scholar 
    Damtew, A., Birhane, E., Messier, C., Paquette, A. & Muys, B. Shading and species diversity act as safety nets for seedling survival and vitality of native trees in dryland forests: implications for restoration. Forest Ecol. Manag. 552, 121559 (2024).Article 

    Google Scholar 
    Williams, L. J., Cavender-Bares, J., Paquette, A., Messier, C. & Reich, P. B. Light mediates the relationship between community diversity and trait plasticity in functionally and phylogenetically diverse tree mixtures. J. Ecol. 108, 1617–1634 (2020).Article 

    Google Scholar 
    Luo, S., Schmid, B., De Deyn, G. B. & Yu, S. Soil microbes promote complementarity effects among co-existing trees through soil nitrogen partitioning. Funct. Ecol. 32, 1879–1889 (2018).Article 

    Google Scholar 
    Soudzilovskaia, N. A. et al. Global mycorrhizal plant distribution linked to terrestrial carbon stocks. Nat. Commun. 10, 5077 (2019).Article 

    Google Scholar 
    Averill, C., Bhatnagar, J. M., Dietze, M. C., Pearse, W. D. & Kivlin, S. N. Global imprint of mycorrhizal fungi on whole-plant nutrient economics. Proc. Natl Acad. Sci. USA 116, 23163–23168 (2019).Article 
    CAS 

    Google Scholar 
    Deng, M. et al. Tree mycorrhizal association types control biodiversity-productivity relationship in a subtropical forest. Sci. Adv. 9, eadd4468 (2023).Article 

    Google Scholar 
    Luo, S. et al. Mycorrhizal associations modify tree diversity−productivity relationships across experimental tree plantations. New Phytol. 243, 1205–1219 (2024).Article 

    Google Scholar 
    Dietrich, P. et al. Tree diversity effects on productivity depend on mycorrhizae and life strategies in a temperate forest experiment. Ecology 104, e3896 (2023).Article 

    Google Scholar 
    Sachsenmaier, L. et al. Forest growth resistance and resilience to the 2018–2020 drought depend on tree diversity and mycorrhizal type. J. Ecol. 112, 1787–1803 (2024).Article 

    Google Scholar 
    Loreau, M. Does functional redundancy exist? Oikos 104, 606–611 (2004).Article 

    Google Scholar 
    Schmid, B. et al. in Biodiversity, Ecosystem Functioning, and Human Wellbeing: An Ecological and Economic Perspective (eds Shahid N. et al.) 14–29 (Oxford University Press, 2009).Hector, A. & Bagchi, R. Biodiversity and ecosystem multifunctionality. Nature 448, 188–190 (2007).Article 
    CAS 

    Google Scholar 
    Reich, P. B. et al. Impacts of biodiversity loss escalate through time as redundancy fades. Science 336, 589–592 (2012).Article 
    CAS 

    Google Scholar 
    Wagg, C. et al. Biodiversity-stability relationships strengthen over time in a long-term grassland experiment. Nat. Commun. 13, 7752 (2022).Article 
    CAS 

    Google Scholar 
    Zheng, L. et al. Effects of plant diversity on productivity strengthen over time due to trait-dependent shifts in species overyielding. Nat. Commun. 15, 2078 (2024).Article 
    CAS 

    Google Scholar 
    Kunz, M. et al. Neighbour species richness and local structural variability modulate aboveground allocation patterns and crown morphology of individual trees. Ecol. Lett. 22, 2130–2140 (2019).Article 

    Google Scholar 
    Shovon, T. A., Kang, S., Scherer-Lorenzen, M. & Nock, C. A. Changes in the direction of the diversity-productivity relationship over 15 years of stand development in a planted temperate forest. J. Ecol. 110, 1125–1137 (2022).Article 

    Google Scholar 
    Jucker, T. et al. Good things take time — diversity effects on tree growth shift from negative to positive during stand development in boreal forests. J. Ecol. 108, 2198–2211 (2020).Article 

    Google Scholar 
    Meyer, S. T. et al. Effects of biodiversity strengthen over time as ecosystem functioning declines at low and increases at high biodiversity. Ecosphere 7, e01619 (2016).Article 

    Google Scholar 
    Guerrero-Ramírez, N. R. et al. Diversity-dependent temporal divergence of ecosystem functioning in experimental ecosystems. Nat. Ecol. Evol. 1, 1639–1642 (2017).Article 

    Google Scholar 
    Brassard, B. W., Chen, H. Y. H., Bergeron, Y. & Paré, D. Differences in fine root productivity between mixed- and single-species stands. Funct. Ecol. 25, 238–246 (2011).Article 

    Google Scholar 
    Urgoiti, J., Messier, C., Keeton, W. S., Belluau, M. & Paquette, A. Functional diversity and identity influence the self-thinning process in young forest communities. J. Ecol. 111, 2010–2022 (2023).Article 

    Google Scholar 
    Chen, X., Chen, H. Y. H., Searle, E. B., Chen, C. & Reich, P. B. Negative to positive shifts in diversity effects on soil nitrogen over time. Nat. Sustain. 4, 225–232 (2021).Article 

    Google Scholar 
    Thakur, M. P. et al. Plant–soil feedbacks and temporal dynamics of plant diversity–productivity relationships. Trends Ecol. Evol. 36, 651–661 (2021).Article 

    Google Scholar 
    Eisenhauer, N., Reich, P. B. & Scheu, S. Increasing plant diversity effects on productivity with time due to delayed soil biota effects on plants. Basic Appl. Ecol. 13, 571–578 (2012).Article 

    Google Scholar 
    Zuppinger-Dingley, D. et al. Selection for niche differentiation in plant communities increases biodiversity effects. Nature 515, 108–111 (2014).Article 
    CAS 

    Google Scholar 
    Ang, C. C. et al. Genetic diversity of two tropical tree species of the Dipterocarpaceae following logging and restoration in Borneo: high genetic diversity in plots with high species diversity. Plant Ecol. Divers. 9, 459–469 (2016).Article 

    Google Scholar 
    Ma, Z. & Chen, H. Y. H. Effects of species diversity on fine root productivity in diverse ecosystems: a global meta-analysis. Glob. Ecol. Biogeogr. 25, 1387–1396 (2016).Article 

    Google Scholar 
    Hisano, M., Chen, H. Y. H., Searle, E. B. & Reich, P. B. Species-rich boreal forests grew more and suffered less mortality than species-poor forests under the environmental change of the past half-century. Ecol. Lett. 22, 999–1008 (2019).Article 

    Google Scholar 
    Taylor, A. R., Gao, B. L. & Chen, H. Y. H. The effect of species diversity on tree growth varies during forest succession in the boreal forest of central Canada. Forest Ecol. Manag. 455, 117641 (2020).Article 

    Google Scholar 
    Schmid, B. et al. Removing subordinate species in a biodiversity experiment to mimic observational field studies. Grassl. Res. 1, 53–62 (2022).Article 
    CAS 

    Google Scholar 
    Hong, P. et al. Biodiversity promotes ecosystem functioning despite environmental change. Ecol. Lett. 25, 555–569 (2022).Article 

    Google Scholar 
    He, J.-S., Bazzaz, F. A. & Schmid, B. Interactive effects of diversity, nutrients and elevated CO2 on experimental plant communities. Oikos 97, 337–348 (2002).Article 
    CAS 

    Google Scholar 
    Fridley, J. D. Diversity effects on production in different light and fertility environments: an experiment with communities of annual plants. J. Ecol. 91, 396–406 (2003).Article 

    Google Scholar 
    Jousset, A., Schmid, B., Scheu, S. & Eisenhauer, N. Genotypic richness and dissimilarity opposingly affect ecosystem functioning. Ecol. Lett. 14, 537–545 (2011).Article 
    CAS 

    Google Scholar 
    Bertness, M. D. & Callaway, R. Positive interactions in communities. Trends Ecol. Evol. 9, 191–193 (1994).Article 
    CAS 

    Google Scholar 
    Gamfeldt, L. et al. Scaling-up the biodiversity-ecosystem functioning relationship: the effect of environmental heterogeneity on transgressive overyielding. Oikos 2023, e09652 (2023).Article 

    Google Scholar 
    Ratcliffe, S. et al. Biodiversity and ecosystem functioning relations in European forests depend on environmental context. Ecol. Lett. 20, 1414–1426 (2017).Article 

    Google Scholar 
    Yan, G. et al. Climate and mycorrhizae mediate the relationship of tree species diversity and carbon stocks in subtropical forests. J. Ecol. 110, 2462–2474 (2022).Article 

    Google Scholar 
    Paquette, A., Vayreda, J., Coll, L., Messier, C. & Retana, J. Climate change could negate positive tree diversity effects on forest productivity: a study across five climate types in Spain and Canada. Ecosystems 21, 960–970 (2018).Article 

    Google Scholar 
    Dhiedt, E., Verheyen, K., De Smedt, P., Ponette, Q. & Baeten, L. Early tree diversity and composition effects on topsoil chemistry in young forest plantations depend on site context. Ecosystems 24, 1638–1653 (2021).Article 
    CAS 

    Google Scholar 
    Belluau, M., Vitali, V., Parker, W. C., Paquette, A. & Messier, C. Overyielding in young tree communities does not support the stress-gradient hypothesis and is favoured by functional diversity and higher water availability. J. Ecol. 109, 1790–1803 (2021).Article 
    CAS 

    Google Scholar 
    Zheng, L. et al. Neighbourhood diversity increases tree growth in experimental forests more in wetter climates but not in wetter years. Nat. Ecol. Evol. 9, 1812–1824 (2025).Article 

    Google Scholar 
    Eisenhauer, N. et al. in Advances in Ecological Research Vol. 61, Ch. 1 (eds Eisenhauer, N., Bohan, D. A. & Dumbrell, A. J.) 1–54 (Academic Press, 2019).Schuldt, A. et al. Multiple plant diversity components drive consumer communities across ecosystems. Nat. Commun. 10, 1460 (2019).Article 

    Google Scholar 
    Zemp, D. C. et al. Tree islands enhance biodiversity and functioning in oil palm landscapes. Nature 618, 316–321 (2023).Article 
    CAS 

    Google Scholar 
    Rutten, G. et al. More diverse tree communities promote foliar fungal pathogen diversity, but decrease infestation rates per tree species, in a subtropical biodiversity experiment. J. Ecol. 109, 2068–2080 (2021).Article 

    Google Scholar 
    Butz, E. M., Schmitt, L. M., Parker, J. D. & Burghardt, K. T. Positive tree diversity effects on arboreal spider abundance are tied to canopy cover in a forest experiment. Ecology 104, e4116 (2023).Article 

    Google Scholar 
    Chen, J. T. et al. Functional and phylogenetic relationships link predators to plant diversity via trophic and non-trophic pathways. Proc. Biol Sci. 290, 20221658 (2023).
    Google Scholar 
    Guo, P. F. et al. Tree diversity promotes predatory wasps and parasitoids but not pollinator bees in a subtropical experimental forest. Basic Appl. Ecol. 53, 134–142 (2021).Article 

    Google Scholar 
    Staab, M. & Schuldt, A. The influence of tree diversity on natural enemies — a review of the “Enemies” Hypothesis in Forests. Curr. Forestry Rep. 6, 243–259 (2020).Article 

    Google Scholar 
    Wang, M. Q. et al. Phylogenetic relatedness, functional traits, and spatial scale determine herbivore co-occurrence in a subtropical forest. Ecol. Monogr. 92, e01492 (2022).Article 

    Google Scholar 
    Grossman, J. J. et al. Synthesis and future research directions linking tree diversity to growth, survival, and damage in a global network of tree diversity experiments. Environ. Exp. Botany 152, 68–89 (2018).Article 

    Google Scholar 
    Abdala-Roberts, L. et al. Effects of tree species diversity and conspecific seedling density on insect herbivory and pathogen infection on big-leaf mahogany seedlings. Oikos https://doi.org/10.1111/oik.10093 (2023).Article 

    Google Scholar 
    Jactel, H., Moreira, X. & Castagneyrol, B. Tree diversity and forest resistance to insect pests: patterns, mechanisms, and prospects. Annu. Rev. Entomol. 66, 277–296 (2021).Article 
    CAS 

    Google Scholar 
    Stemmelen, A., Jactel, H., Brockerhoff, E. & Castagneyrol, B. Meta-analysis of tree diversity effects on the abundance, diversity and activity of herbivores’ enemies. Basic Appl. Ecol. 58, 130–138 (2022).Article 

    Google Scholar 
    Vázquez-González, C. et al. Tree diversity enhances predation by birds but not by arthropods across climate gradients. Ecol. Lett. 27, e14427 (2024).Article 

    Google Scholar 
    Garau, G. et al. Effect of monospecific and mixed Mediterranean tree plantations on soil microbial community and biochemical functioning. Appl. Soil Ecol. 140, 78–88 (2019).Article 

    Google Scholar 
    Strukelj, M. et al. Tree species richness and water availability interact to affect soil microbial processes. Soil Biol. Biochem. 155, 108180 (2021).Article 
    CAS 

    Google Scholar 
    Beugnon, R. et al. Tree diversity effects on litter decomposition are mediated by litterfall and microbial processes. Oikos https://doi.org/10.1111/oik.09751 (2023).Article 

    Google Scholar 
    Cesarz, S. et al. Tree diversity effects on soil microbial biomass and respiration are context dependent across forest diversity experiments. Glob. Ecol. Biogeogr. 31, 872–885 (2022).Article 

    Google Scholar 
    Rivest, M., Whalen, J. K. & Rivest, D. Tree diversity is not always a strong driver of soil microbial diversity: a 7-yr-old diversity experiment with trees. Ecosphere 10, e02685 (2019).Article 

    Google Scholar 
    Tao, S. Q., Veen, G. F., Zhang, N. L., Yu, T. H. & Qu, L. Y. Tree and shrub richness modifies subtropical tree productivity by regulating the diversity and community composition of soil bacteria and archaea. Microbiome 11, 261 (2023).Article 
    CAS 

    Google Scholar 
    Yang, X. et al. Different assembly mechanisms of leaf epiphytic and endophytic bacterial communities underlie their higher diversity in more diverse forests. J. Ecol. 111, 970–981 (2023).Article 

    Google Scholar 
    Setiawan, N. N. et al. Does neighbourhood tree diversity affect the crown arthropod community in saplings? Biodivers. Conserv. 25, 169–185 (2016).Article 

    Google Scholar 
    Zhang, S. et al. Non-random tree species loss shifts soil fungal communities. J. Ecol. 113, 1239–1255 (2025).Article 
    CAS 

    Google Scholar 
    Berthelot, S. et al. Exotic tree species have consistently lower herbivore load in a cross-Atlantic tree biodiversity experiment. Ecology 104, e4070 (2023).Article 

    Google Scholar 
    Castagneyrol, B., Jactel, H. & Moreira, X. Anti-herbivore defences and insect herbivory: interactive effects of drought and tree neighbours. J. Ecol. 106, 2043–2057 (2018).Article 
    CAS 

    Google Scholar 
    Field, E. et al. Associational resistance to both insect and pathogen damage in mixed forests is modulated by tree neighbour identity and drought. J. Ecol. 108, 1511–1522 (2020).Article 
    CAS 

    Google Scholar 
    Li, Y. et al. The tree growth–herbivory relationship depends on functional traits across forest biodiversity experiments. Nat. Ecol. Evol. 9, 2014–2024 (2025).Article 

    Google Scholar 
    Poeydebat, C. et al. Climate affects neighbour-induced changes in leaf chemical defences and tree diversity-herbivory relationships. Funct. Ecol. 35, 67–81 (2021).Article 
    CAS 

    Google Scholar 
    Schmitz, O. J. Effects of predator hunting mode on grassland ecosystem function. Science 319, 952–954 (2008).Article 
    CAS 

    Google Scholar 
    Haddad, N. M. et al. Plant species loss decreases arthropod diversity and shifts trophic structure. Ecol. Lett. 12, 1029–1039 (2009).Article 

    Google Scholar 
    Scherber, C. et al. Bottom-up effects of plant diversity on multitrophic interactions in a biodiversity experiment. Nature 468, 553–556 (2010).Article 
    CAS 

    Google Scholar 
    Weisser, W. W. et al. Biodiversity effects on ecosystem functioning in a 15-year grassland experiment: patterns, mechanisms, and open questions. Basic Appl. Ecol. 23, 1–73 (2017).Article 

    Google Scholar 
    Scherber, C. Convergent patterns in multitrophic biodiversity effects on yield across ecosystems. Sci. China Life Sci. 66, 2697–2699 (2023).Article 

    Google Scholar 
    Brezzi, M., Schmid, B., Niklaus, P. A. & Schuldt, A. Tree diversity increases levels of herbivore damage in a subtropical forest canopy: evidence for dietary mixing by arthropods? J. Plant Ecol. 10, 13–27 (2017).Article 

    Google Scholar 
    van der Plas, F. Biodiversity and ecosystem functioning in naturally assembled communities. Biol. Rev. 94, 1220–1245 (2019).Article 

    Google Scholar 
    Kremer, K., Jonsson, B.-G., Dutta, T., Tavares, M. F. & Bauhus, J. Single- vs mixed-species plantations: a systematic review on the effects on biodiversity. Biol. Conserv. 307, 111182 (2025).Article 

    Google Scholar 
    Gottschall, F. et al. Spatiotemporal dynamics of abiotic and biotic properties explain biodiversity–ecosystem-functioning relationships. Ecol. Monogr. 92, e01490 (2022).Article 

    Google Scholar 
    May-Uc, Y., Nell, C. S., Parra-Tabla, V., Augusto, J. & Abdala-Roberts, L. Tree diversity effects through a temporal lens: implications for the abundance, diversity and stability of foraging birds. J. Anim. Ecol. 89, 1775–1787 (2020).Article 

    Google Scholar 
    Wang, M.-Q. et al. Tree diversity, tree growth, and microclimate independently structure Lepidoptera herbivore community stability. Ecol. Monogr. 95, e70026 (2025).Article 

    Google Scholar 
    Huang, Y. et al. Effects of enemy exclusion on biodiversity–productivity relationships in a subtropical forest experiment. J. Ecol. 110, 2167–2178 (2022).Article 

    Google Scholar 
    Fornoff, F., Klein, A.-M., Blüthgen, N. & Staab, M. Tree diversity increases robustness of multi-trophic interactions. Proc. Biol. Sci. 286, 20182399 (2019).
    Google Scholar 
    Potapov, A. M. et al. Rainforest transformation reallocates energy from green to brown food webs. Nature 627, 116–122 (2024).Article 
    CAS 

    Google Scholar 
    Wang, S., Brose, U. & Gravel, D. Intraguild predation enhances biodiversity and functioning in complex food webs. Ecology 100, e02616 (2019).Article 

    Google Scholar 
    Hennessy, A. B., Anderson, R. M., Mitchell, N., Mooney, K. A. & Singer, M. S. Parasitoid avoidance of intraguild predation drives enemy complementarity in a multi-trophic ecological network. Ecology 106, e4483 (2025).Article 

    Google Scholar 
    Eisenhauer, N. et al. Ecosystem consequences of invertebrate decline. Curr. Biol. 33, 4538–4547 (2023).Article 
    CAS 

    Google Scholar 
    Albert, G., Gauzens, B., Loreau, M., Wang, S. & Brose, U. The hidden role of multi-trophic interactions in driving diversity–productivity relationships. Ecol. Lett. 25, 405–415 (2022).Article 

    Google Scholar 
    Albert, G. et al. Animal and plant space-use drive plant diversity–productivity relationships. Ecol. Lett. 26, 1792–1802 (2023).Article 

    Google Scholar 
    Lepš, J. What do the biodiversity experiments tell us about consequences of plant species loss in the real world? Basic Appl. Ecol. 5, 529–534 (2004).Article 

    Google Scholar 
    Jochum, M. et al. The results of biodiversity–ecosystem functioning experiments are realistic. Nat. Ecol. Evol. 4, 1485–1494 (2020).Article 

    Google Scholar 
    Baruffol, M. et al. Biodiversity promotes tree growth during succession in subtropical forest. PLoS ONE 8, e81246 (2013).Article 

    Google Scholar 
    Liu, X. et al. Species richness, functional traits and climate interactively affect tree survival in a large forest biodiversity experiment. J. Ecol. 110, 2522–2531 (2022).Article 

    Google Scholar 
    Marquard, E. et al. Positive biodiversity-productivity relationship due to increased plant density. J. Ecol. 97, 696–704 (2009).Article 

    Google Scholar 
    Kambach, S. et al. How do trees respond to species mixing in experimental compared to observational studies? Ecol. Evol. 9, 11254–11265 (2019).Article 

    Google Scholar 
    Hagan, J. G., Vanschoenwinkel, B. & Gamfeldt, L. We should not necessarily expect positive relationships between biodiversity and ecosystem functioning in observational field data. Ecol. Lett. 24, 2537–2548 (2021).Article 

    Google Scholar 
    Pärtel, M. et al. Global impoverishment of natural vegetation revealed by dark diversity. Nature 641, 917–924 (2025).Article 

    Google Scholar 
    García-Valdés, R., Bugmann, H. & Morin, X. Climate change-driven extinctions of tree species affect forest functioning more than random extinctions. Diversity Distrib. 24, 906–918 (2018).Article 

    Google Scholar 
    Oehri, J., Schmid, B., Schaepman-Strub, G. & Niklaus, P. A. Terrestrial land-cover type richness is positively linked to landscape-level functioning. Nat. Commun. 11, 154 (2020).Article 
    CAS 

    Google Scholar 
    Messier, C. et al. For the sake of resilience and multifunctionality, let’s diversify planted forests! Conserv. Lett. 15, e12829 (2022).Article 

    Google Scholar 
    Smith, P. et al. How do we best synergize climate mitigation actions to co-benefit biodiversity? Glob. Change Biol. 28, 2555–2577 (2022).Article 
    CAS 

    Google Scholar 
    Bongers, F. J. et al. Genetic richness affects trait variation but not community productivity in a tree diversity experiment. New Phytol. 227, 744–756 (2020).Article 

    Google Scholar 
    Allan, E., Penone, C., Schmid, B., Godoy, O. & Pichon, N. A. When can we expect negative effects of plant diversity on community biomass? J. Ecol. 113, 1955–1969 (2025).Article 

    Google Scholar 
    Di Maurizio, V., Searle, E. & Paquette, A. It takes a village to grow a tree: most tree species benefit from dissimilar neighbors. Ecol. Evol. 13, e10804 (2023).Article 

    Google Scholar 
    Beugnon, R. et al. Improving forest ecosystem functions by optimizing tree species spatial arrangement. Nat. Commun. 16, 6286 (2025).Article 
    CAS 

    Google Scholar 
    Warner, E. et al. Young mixed planted forests store more carbon than monocultures — a meta-analysis. Front. Forests Glob. Change 6, 1226514 (2023).Article 

    Google Scholar 
    Paquette, A. & Messier, C. The role of plantations in managing the world’s forests in the Anthropocene. Front. Ecol. Environ. 8, 27–34 (2010).Article 

    Google Scholar 
    Urgoiti Otazua, J. & Paquette, A. in Dynamics, Silviculture and Management of Mixed Forests (eds Bravo-Oviedo, A., Pretzsch, H. & del Río, M.) 319–341 (Springer International Publishing, 2018).Wang, S. P. et al. Towards mechanistic integration of the causes and consequences of biodiversity. Trends Ecol. Evol. 39, 689–700 (2024).Article 

    Google Scholar 
    Chen, X. L. et al. Tree diversity increases decadal forest soil carbon and nitrogen accrual. Nature 618, 94–101 (2023).Article 
    CAS 

    Google Scholar 
    Handa, I. T. et al. Consequences of biodiversity loss for litter decomposition across biomes. Nature 509, 218–221 (2014).Article 
    CAS 

    Google Scholar 
    Scherer-Lorenzen, M., Luis Bonilla, J. & Potvin, C. Tree species richness affects litter production and decomposition rates in a tropical biodiversity experiment. Oikos 116, 2108–2124 (2007).Article 

    Google Scholar 
    Li, Y. et al. Early positive effects of tree species richness on soil organic carbon accumulation in a large-scale forest biodiversity experiment. J. Plant Ecol. 12, 882–893 (2019).Article 

    Google Scholar 
    Schnabel, F. et al. Tree diversity increases carbon stocks and fluxes above — but not belowground in a tropical forest experiment. Glob. Change Biol. 31, e70089 (2025).Article 
    CAS 

    Google Scholar 
    Martin-Guay, M.-O., Paquette, A., Reich, P. B. & Messier, C. Implications of contrasted above- and below-ground biomass responses in a diversity experiment with trees. J. Ecol. 108, 405–414 (2020).Article 

    Google Scholar 
    Isbell, F. et al. Biodiversity increases the resistance of ecosystem productivity to climate extremes. Nature 526, 574–577 (2015).Article 
    CAS 

    Google Scholar 
    Eisenhauer, N. et al. Plant diversity effects on soil food webs are stronger than those of elevated CO2 and N deposition in a long-term grassland experiment. Proc. Natl Acad. Sci. USA 110, 6889–6894 (2013).Article 
    CAS 

    Google Scholar 
    Yang, B. et al. Soil fungi promote biodiversity–productivity relationships in experimental communities of young trees. Ecosystems 25, 858–871 (2022).Article 
    CAS 

    Google Scholar 
    Riedel, J., Dorn, S., Plath, M., Potvin, C. & Mody, K. Time matters: temporally changing effects of planting schemes and insecticide treatment on native timber tree performance on former pasture. Forest Ecol. Manag. 297, 49–56 (2013).Article 

    Google Scholar 
    Schmid, B. in Defining Agroecology (eds Dormann, C. F. et al.) 143–156 (tredition.com, 2023).Lienau, J. R., Duguid, M. C. & Schmitz, O. J. Ground beetle trophic interactions alter available nitrogen in forest soil. Oikos https://doi.org/10.1111/oik.10638 (2025).Article 

    Google Scholar 
    Sivault, E. et al. Insectivorous birds and bats outperform ants in the top-down regulation of arthropods across strata of a Japanese temperate forest. J. Anim. Ecol. 93, 1622–1638 (2024).Article 

    Google Scholar 
    Cook-Patton, S. C., LaForgia, M. & Parker, J. D. Positive interactions between herbivores and plant diversity shape forest regeneration. Proc. Biol. Sci. 281, 20140261 (2014).
    Google Scholar 
    Singer, M. S. et al. Herbivore diet breadth mediates the cascading effects of carnivores in food webs. Proc. Natl Acad. Sci. USA 111, 9521–9526 (2014).Article 
    CAS 

    Google Scholar 
    Yeeles, P., Lach, L., Hobbs, R. J. & Didham, R. K. Functional redundancy compensates for decline of dominant ant species. Nat. Ecol. Evol. 9, 779–788 (2025).Article 

    Google Scholar 
    Liang, M. et al. Unifying spatial scaling laws of biodiversity and ecosystem stability. Science 387, eadl2373 (2025).Article 
    CAS 

    Google Scholar 
    Siegel, K. & Dee, L. E. Foundations and future directions for causal inference in ecological research. Ecol. Lett. 28, e70053 (2025).Article 

    Google Scholar 
    Imbens, G. W. & Rubin, D. B. Causal Inference for Statistics, Social, and Biomedical Sciences: An Introduction (Cambridge University Press, 2015).Damtew, A., Birhane, E., Messier, C., Paquette, A. & Muys, B. Shading and selection effect-mediated species mixing enhance the growth of native trees in dry tropical forests. Oecologia 207, 75 (2025).Article 

    Google Scholar 
    Van de Peer, T. et al. Tree seedling vitality improves with functional diversity in a Mediterranean common garden experiment. Forest Ecol. Manag. 409, 614–633 (2018).Article 

    Google Scholar 
    Schuldt, A., Fornoff, F., Bruelheide, H., Klein, A.-M. & Staab, M. Tree species richness attenuates the positive relationship between mutualistic ant–hemipteran interactions and leaf chewer herbivory. Proc. Biol. Sci. 284, 20171489 (2017).
    Google Scholar 
    Snedecor, G. W. & Cochran, W. G. Statistical Methods, 8th edn (Iowa State University Press, 1989).Schmid, B., Polasek, W., Weiner, J., Krause, A. & Stoll, P. Modeling of discontinuous relationships in biology with censored regression. Am. Naturalist 143, 494–507 (1994).Article 

    Google Scholar 
    Schmid, B. The species richness-productivity controversy. Trends Ecol. Evol. 17, 113–114 (2002).Article 

    Google Scholar 
    Naeem, S., Thompson, L. J., Lawler, S. P., Lawton, J. H. & Woodfin, R. M. Declining biodiversity can alter the performance of ecosystems. Nature 368, 734–737 (1994).Article 

    Google Scholar 
    Loreau, M. et al. Biodiversity as insurance: from concept to measurement and application. Biol. Rev. 96, 2333–2354 (2021).Article 

    Google Scholar 
    Grace, J. B. et al. Causal effects versus causal mechanisms: two traditions with different requirements and contributions towards causal understanding. Ecol. Lett. 28, e70029 (2025).Article 

    Google Scholar 
    Pearl, J. Causality 2 edn (Cambridge University Press, 2009).Andraczek, K. et al. Weak reciprocal relationships between productivity and plant biodiversity in managed grasslands. J. Ecol. 112, 2359–2373 (2024).Article 

    Google Scholar 
    Runge, J. Modern causal inference approaches to investigate biodiversity-ecosystem functioning relationships. Nat. Commun. 14, 1917 (2023).Article 
    CAS 

    Google Scholar 
    Barry, K. E. et al. The future of complementarity: disentangling causes from consequences. Trends Ecol. Evol. 34, 167–180 (2019).Article 

    Google Scholar 
    Grime, J. P. Benefits of plant diversity to ecosystems: immediate, filter and founder effects. J. Ecol. 86, 902–910 (1998).Article 

    Google Scholar 
    Download referencesAcknowledgementsWe thank S. Li, Y. Li, C. Chen and S. Zhang for their help in collecting the experiment information in Supplementary Table 1 and improving the figures. X.L. was supported by the National Key Research Development Program of China (2022YFF0802300), the National Natural Science Foundation of China (32525042 and 32222055), and the Youth Innovation Promotion Association CAS (2023019). J.C.-B. was supported by the ASCEND Biology Integration Institute, NSF DBI (2021898) and Cedar Creek Long-Term Ecological Research, NSF DEB (1831944). A.S. was supported by the German Research Foundation DFG (452861007/FOR 5281). B.S. was supported by the NOMIS Foundation, the Presidential International Fellowship Initiative (PIFI) from the Chinese Academy of Sciences and the University Research Priority Program on Global Change and Biodiversity of the University of Zurich.Author informationAuthors and AffiliationsKey Laboratory of Vegetation and Environmental Change, Institute of Botany, Chinese Academy of Sciences, Beijing, ChinaXiaojuan Liu & Keping MaChina National Botanical Garden, Beijing, ChinaXiaojuan LiuUniversity of Chinese Academy of Sciences, Beijing, ChinaXiaojuan LiuForest Nature Conservation, University of Göttingen, Göttingen, GermanyAndreas SchuldtDepartment of Organismic and Evolutionary Biology, Harvard University, Cambridge, MA, USAJeannine Cavender-BaresCentre for Forest Research, Université du Québec à Montréal, Montréal, Quebec, CanadaAlain PaquetteRemote Sensing Laboratories, Department of Geography, University of Zurich, Zurich, SwitzerlandBernhard SchmidAuthorsXiaojuan LiuView author publicationsSearch author on:PubMed Google ScholarAndreas SchuldtView author publicationsSearch author on:PubMed Google ScholarJeannine Cavender-BaresView author publicationsSearch author on:PubMed Google ScholarAlain PaquetteView author publicationsSearch author on:PubMed Google ScholarBernhard SchmidView author publicationsSearch author on:PubMed Google ScholarKeping MaView author publicationsSearch author on:PubMed Google ScholarContributionsX.L. and K.M. conceived the idea. X.L. A.S., J.C.-B., A.P., B.S. and K.M. together wrote the review.Corresponding authorCorrespondence to
    Keping Ma.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Peer review

    Peer review information
    Nature Reviews Biodiversity thanks Santiago Soliveres, Marta Monfort and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

    Additional informationPublisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Related linksUnited Nations Convention on Biological Diversity Global Biodiversity Targets: https://www.cbd.int/gbf/targetsSupplementary informationSupplementary informationGlossaryFacilitation
    A species in a mixture benefits from the presence of other species that change the abiotic or biotic environment.
    Functional diversity
    Variation in function among individuals or species within a circumscribed space, often calculated for multiple traits, using a variety of metrics, which can be abundance-weighted.
    Multi-functionality
    Measures combining multiple ecosystem functions into a single value, based on averages or on the number of functions reaching a minimal level.
    Multi-trophic diversity
    The species diversity across multiple groups of organisms belonging to different trophic levels, often expressed as the average across the standardized (that is, relative) species richness values of each group of organisms.
    Phylogenetic diversity
    Evolutionary divergence among individuals or species within a circumscribed space, calculated from a phylogeny using a variety of metrics, which can be abundance-weighted.
    Resilience
    The ability of ecosystem properties to return to a pre-disturbance condition after a disturbance.
    Resistance
    The ability of individuals, species or communities to persist and maintain ecosystem functions despite exposure to a stress or disturbance.
    Stability
    The capacity of an ecosystem to maintain its structure and function over time, despite disturbances and environmental fluctuations.
    Rights and permissionsSpringer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.Reprints and permissionsAbout this articleCite this articleLiu, X., Schuldt, A., Cavender-Bares, J. et al. Ecological insights from three decades of forest biodiversity experiments.
    Nat. Rev. Biodivers. (2026). https://doi.org/10.1038/s44358-025-00112-2Download citationAccepted: 06 November 2025Published: 02 January 2026Version of record: 02 January 2026DOI: https://doi.org/10.1038/s44358-025-00112-2Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Fungal parasites infecting N2-fixing cyanobacteria reshape carbon and N2 fixation and trophic transfer

    AbstractFungal parasites are associated with bloom-forming algae, yet their impact on N2 fixation and the fate of newly fixed nitrogen during cyanobacterial blooms is poorly understood. We report infections on the ecologically important N2-fixing cyanobacterium Dolichospermum (formerly Anabaena) in the Baltic Sea. Using single-cell isotope probing, microscopy, and biogeochemical analyses, we examine how infections affect carbon and N2 fixation and elemental transfer within a natural community. Fungal sporangia infect up to 80% of filaments, mostly targeting storage cells (akinetes, 82% prevalence) and N2-fixing cells (heterocytes, 44%), but rarely vegetative cells (5%). Infections at akinete–heterocyte junctions extract 4- and 10-fold more carbon and nitrogen than those on vegetative cells, reducing host storage by 28% and 56%. Overall, 22% of newly fixed nitrogen is transferred to fungi, comparable to heterotrophic bacteria. Infections also occur in Nodularia and Aphanizomenon, suggesting fungi-like parasitism broadly affects bloom dynamics and the fate of new nitrogen.

    Data availability

    Sequence data have been deposited in ENA (European Nucleotide Archive) under project accession no. PRJEB96922. Accession numbers used to construct the phylogenetic tree in Supplementary Fig. S5 are listed in Supplementary Table S5, as accessed on the NCBI database. The raw mass spectrometer output can be found in Supplementary Dataset 1. Source data are provided with this paper.
    ReferencesFalkowski, P. The power of plankton. Nature 483, 17–20 (2012).
    Google Scholar 
    Chassot, E. et al. Global marine primary production constrains fisheries catches. Ecol. Lett. 13, 495–505 (2010).
    Google Scholar 
    Stock, C. A. et al. Reconciling fisheries catch and ocean productivity. Proc. Natl. Acad. Sci. USA 114, E1441–E1449 (2017).
    Google Scholar 
    Berdalet, E. et al. Marine harmful algal blooms, human health and wellbeing: challenges and opportunities in the 21st century. J. Mar. Biol. Assoc. UK. 2015, 1–62 (2015).Wurtsbaugh, W. A., Paerl, H. W. & Dodds, W. K. Nutrients, eutrophication and harmful algal blooms along the freshwater to marine continuum. Wiley Interdiscip. Rev. Water 6, e1373 (2019).
    Google Scholar 
    Paerl, H. W. & Huisman, J. Blooms like it hot. Science 320, 57–58 (2008).
    Google Scholar 
    Dai, Y. et al. Coastal phytoplankton blooms expand and intensify in the 21st century. Nature 615, 280–284 (2023).
    Google Scholar 
    Kahru, M. & Elmgren, R. Multidecadal time series of satellite-detected accumulations of cyanobacteria in the Baltic Sea. Biogeosciences 11, 3619–3633 (2014).
    Google Scholar 
    Finni, T., Kononen, K., Olsonen, R. & Wallström, K. The history of cyanobacterial blooms in the Baltic Sea. Ambio 30, 172–178 (2001).
    Google Scholar 
    HELCOM & Earth, B. Climate Change in the Baltic Sea 2021 Fact Sheet: Baltic Sea Environmental Proceedings No.180 (Helsinki Commission – HELCOM, 2021).Larsson, U., Hajdu, S., Walve, J. & Elmgren, R. Baltic Sea nitrogen fixation estimated from the summer increase in upper mixed layer total nitrogen. Limnol. Oceanogr. 46, 811–820 (2001).
    Google Scholar 
    Rolff, C., Almesjö, L. & Elmgren, R. Nitrogen fixation and abundance of the diazotrophic cyanobacterium Aphanizomenon sp. in the Baltic Proper. Mar. Ecol. Prog. Ser. 332, 107–118 (2007).
    Google Scholar 
    Funkey, C. P. et al. Hypoxia sustains cyanobacteria blooms in the Baltic Sea. Environ. Sci. Technol. 48, 2598–2602 (2014).
    Google Scholar 
    Vahtera, E. et al. Internal ecosystem feedbacks enhance nitrogen-fixing cyanobacteria blooms and complicate management in the Baltic Sea. Ambio 36, 186–194 (2007).
    Google Scholar 
    Worden, A. Z. et al. Rethinking the marine carbon cycle: Factoring in the multifarious lifestyles of microbes. Science 347, 1257594 (2015).
    Google Scholar 
    Paerl, H. Nutrient and other environmental controls of harmful cyanobacterial blooms along the freshwater-marine continuum. Adv. Exp. Med. Biol. 619, 217–237 (2008).
    Google Scholar 
    Jankowiak, J., Hattenrath-Lehmann, T., Kramer, B. J., Ladds, M. & Gobler, C. J. Deciphering the effects of nitrogen, phosphorus, and temperature on cyanobacterial bloom intensification, diversity, and toxicity in western Lake Erie. Limnol. Oceanogr. 64, 1347–1370 (2019).
    Google Scholar 
    Gutiérrez, M. H., Jara, A. M. & Pantoja, S. Fungal parasites infect marine diatoms in the upwelling ecosystem of the Humboldt current system off central Chile. Environ. Microbiol. 18, 1646–1653 (2016).
    Google Scholar 
    Kilias, E. S. et al. Chytrid fungi distribution and co-occurrence with diatoms correlate with sea ice melt in the Arctic Ocean. Commun. Biol. 3, 183 (2020).
    Google Scholar 
    Hassett, B. T. & Gradinger, R. Chytrids dominate arctic marine fungal communities. Environ. Microbiol. 18, 2001–2009 (2016).
    Google Scholar 
    Sime-Ngando, T. Phytoplankton chytridiomycosis: fungal parasites of phytoplankton and their imprints on the food web dynamics. Front. Microbiol. 3, 1–13 (2012).
    Google Scholar 
    Van den Wyngaert, S. et al. Seasonality of parasitic and saprotrophic zoosporic fungi: linking sequence data to ecological traits. ISME J. 16, 2242–2254 (2022).
    Google Scholar 
    Kagami, M., Gurung, T. B., Yoshida, T. & Urabe, J. To sink or to be lysed? Contrasting fate of two large phytoplankton species in Lake Biwa. Limnol. Oceanogr. 51, 2775–2786 (2006).
    Google Scholar 
    Rasconi, S., Niquil, N. & Sime-Ngando, T. Phytoplankton chytridiomycosis: community structure and infectivity of fungal parasites in aquatic ecosystems. Environ. Microbiol. 14, 2151–2170 (2012).
    Google Scholar 
    Gerphagnon, M., Colombet, J., Latour, D. & Sime-Ngando, T. Spatial and temporal changes of parasitic chytrids of cyanobacteria. Sci. Rep. 7, 6056 (2017).
    Google Scholar 
    Klawonn, I. et al. Characterizing the “fungal shunt”: parasitic fungi on diatoms affect carbon flow and bacterial communities in aquatic microbial food webs. Proc. Natl. Acad. Sci. USA 118, e2102225118 (2021).
    Google Scholar 
    Kagami, M., Miki, T. & Takimoto, G. Mycoloop: chytrids in aquatic food webs. Front. Microbiol. 5, 166 (2014).
    Google Scholar 
    Garvetto, A. et al. Chytrid infecting the bloom-forming marine diatom Skeletonema sp.: morphology, phylogeny and distribution of a novel species within the Rhizophydiales. Fungal Biol. 123, 471–480 (2019).
    Google Scholar 
    Fernández-Valero, A. D. et al. Newly identified diversity of Dinomycetaceae (Rhizophydiales, Chytridiomycota), a family of fungal parasites of marine dinoflagellates. Eur. J. Protistol. 93, 126053 (2024).
    Google Scholar 
    McKindles, K. M., Jorge, A. N., McKay, R. M., Davis, T. W. & Bullerjahn, G. S. Isolation and characterization of Rhizophydiales (Chytridiomycota), obligate parasites of Planktothrix agardhii in a Laurentian Great Lakes embayment. Appl. Environ. Microbiol. 87, e02308–e02320 (2021).
    Google Scholar 
    Feng, L. et al. Harmful algal blooms in inland waters. Nat. Rev. Earth Environ. 5, 631–644 (2024).
    Google Scholar 
    Munkes, B., Löptien, U. & Dietze, H. Cyanobacteria blooms in the Baltic Sea: a review of models and facts. Biogeosciences 18, 2347–2378 (2021).
    Google Scholar 
    Kuliński, K. et al. Biogeochemical functioning of the Baltic Sea. Earth Syst. Dyn. 13, 633–685 (2022).
    Google Scholar 
    Karlson, B. et al. Harmful algal blooms and their effects in coastal seas of Northern Europe. Harmful Algae 102, 101989 (2021).
    Google Scholar 
    Reusch, T. B. H. et al. The Baltic Sea as a time machine for the future coastal ocean. Sci. Adv. 4, eaar8195 (2018).
    Google Scholar 
    Takano, K. et al. Fungal infection for cyanobacterium Anabaena smithii by two chytrids in eutrophic region of large reservoir Lake Shumarinai, Hokkaido, Japan. Limnology 9, 213–218 (2008).
    Google Scholar 
    Paterson, R. A. Infestation of Chytridiaceous fungi on phytoplankton in relation to certain environmental factors. Ecology 41, 416–424 (1960).
    Google Scholar 
    Gerphagnon, M., Latour, D., Colombet, J. & Sime-Ngando, T. Fungal parasitism: life cycle, dynamics and impact on cyanobacterial blooms. PLoS ONE 8, e60894 (2013).
    Google Scholar 
    Xu, X., Kasada, M., Grossart, H.-P., Ibelings, B. W. & Van den Wyngaert, S. A newly isolated chytrid fungus specialized in parasitizing heterocysts of the filamentous cyanobacterium Dolichospermum sp. Hydrobiologia 852, 5163–5181 (2025).Canter, H. M. & Lund, J. W. G. Fungal Parasites of the Phytoplankton. II (Studies on British Chytrids, XII). Ann. Bot. 15, 129–156 (1951).Van den Wyngaert, S. et al. Dynamics of zoosporic parasites in summer phytoplankton communities of the Baltic Sea. FEMS Microbiol. Ecol. 101, fiaf081 (2025).Ger, K. A. et al. The interaction between cyanobacteria and zooplankton in a more eutrophic world. Harmful Algae 54, 128–144 (2016).
    Google Scholar 
    Haraldsson, M. et al. Microbial parasites make cyanobacteria blooms less of a trophic dead end than commonly assumed. ISME J. 12, 1008–1020 (2018).
    Google Scholar 
    Frenken, T., Wolinska, J., Tao, Y., Rohrlack, T. & Agha, R. Infection of filamentous phytoplankton by fungal parasites enhances herbivory in pelagic food webs. Limnol. Oceanogr. 65, 2618–2626 (2020).
    Google Scholar 
    Agha, R., Saebelfeld, M., Manthey, C., Rohrlack, T. & Wolinska, J. Chytrid parasitism facilitates trophic transfer between bloom-forming cyanobacteria and zooplankton (Daphnia). Sci. Rep. 6, 35039 (2016).
    Google Scholar 
    Frenken, T. et al. The potential of zooplankton in constraining chytrid epidemics in phytoplankton hosts. Ecology 101, e02900 (2020).
    Google Scholar 
    Kagami, M., Von Elert, E., Ibelings, B. W., De Bruin, A. & Van Donk, E. The parasitic chytrid, Zygorhizidium, facilitates the growth of the cladoceran zooplankter, Daphnia, in cultures of the inedible alga, Asterionella. Proc. Biol. Sci. 274, 1561–1566 (2007).
    Google Scholar 
    Rasconi, S. et al. Parasitic chytrids upgrade and convey primary produced carbon during inedible algae proliferation. Protist 171, 125768 (2020).
    Google Scholar 
    Fay, P. Cell differentiation and pigment composition in Anabaena cylindrica. Arch. fur Mikrobiol. 67, 62–70 (1969).
    Google Scholar 
    Simon, R. D. Sporulation in the filamentous cyanobacterium Anabaena cylindrica. The course of spore formation. Arch. Microbiol. 111, 283–288 (1977).
    Google Scholar 
    Sutherland, J. M., Herdman, M. & Stewart, W. D. P. Akinetes of the cyanobacterium Nostoc PCC 7524: macromolecular composition, structure and control of differentiation. J. Gen. Microbiol. 115, 273–287 (1979).
    Google Scholar 
    Pereira, S. et al. Complexity of cyanobacterial exopolysaccharides: composition, structures, inducing factors and putative genes involved in their biosynthesis and assembly. FEMS Microbiol. Rev. 33, 917–941 (2009).
    Google Scholar 
    Sánchez Barranco, V. et al. Trophic position, elemental ratios and nitrogen transfer in a planktonic host–parasite–consumer food chain including a fungal parasite. Oecologia 194, 541–554 (2020).
    Google Scholar 
    Rasconi, S., Grami, B., Niquil, N., Jobard, M. & Sime-Ngando, T. Parasitic chytrids sustain zooplankton growth during inedible algal bloom. Front. Microbiol. 5, 229 (2014).
    Google Scholar 
    Grasshoff, K., Kremling, K. & Ehrhardt, M. Methods of Seawater Analysis 3rd edn, 159–228 (Wiley-VCH, 1999).Klawonn, I., Dunker, S., Kagami, M., Grossart, H.-P. & Van den Wyngaert, S. Intercomparison of two fluorescent dyes to visualize parasitic fungi (Chytridiomycota) on phytoplankton. Microb. Ecol. 85, 9–23 (2023).
    Google Scholar 
    Olenina, I. et al. Biovolumes and size classes of phytoplankton in the Baltic Sea. HELCOM Balt. Sea Environ. Proc. 106, 144 (2006).
    Google Scholar 
    Wacklin, P., Hoffmann, L. & Komárek, J. Nomenclatural validation of the genetically revised cyanobacterial genus Dolichospermum (RALFS ex BORNET et FLAHAULT) comb. nova. Fottea 9, 59–64 (2009).
    Google Scholar 
    Rueden, C. T. et al. ImageJ2: ImageJ for the next generation of scientific image data. BMC Bioinformatics 18, 529 (2017).
    Google Scholar 
    Nercessian, O., Noyes, E., Kalyuzhnaya, M. G., Lidstrom, M. E. & Chistoserdova, L. Bacterial populations active in metabolism of C1 compounds in the sediment of Lake Washington, a freshwater lake. Appl. Environ. Microbiol. 71, 6885–6899 (2005).
    Google Scholar 
    Vainio, E. J. & Hantula, J. Direct analysis of wood-inhabiting fungi using denaturing gradient gel electrophoresis of amplified ribosomal DNA. Mycol. Res. 104, 927–936 (2000).
    Google Scholar 
    Banos, S. et al. A comprehensive fungi-specific 18S rRNA gene sequence primer toolkit suited for diverse research issues and sequencing platforms. BMC Microbiol 18, 190 (2018).
    Google Scholar 
    Callahan, B. J. et al. DADA2: high-resolution sample inference from Illumina amplicon data. Nat. Methods 13, 581–583 (2016).
    Google Scholar 
    RStudio Team. RStudio: Integrated Development Environment for R https://www.rstudio.com/ (RStudio Team, 2023).Davis, N. M., Proctor, D. M., Holmes, S. P., Relman, D. A. & Callahan, B. J. Simple statistical identification and removal of contaminant sequences in marker-gene and metagenomics data. Microbiome 6, 226 (2018).
    Google Scholar 
    Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genom. Biol. 15, 550 (2014).
    Google Scholar 
    Montoya, J. P., Voss, M., Kähler, P. & Capone, D. G. A simple, high-precision, high-sensitivity tracer assay for N2 fixation. Appl. Environ. Microbiol. 62, 986–993 (1996).
    Google Scholar 
    Pett-Ridge, J. & Weber, P. K. NanoSIP: NanoSIMS applications for microbial biology. Methods Mol. Biol. 2349, 91–136 (2022).Stryhanyuk, H. et al. Calculation of single cell assimilation rates from SIP-nanoSIMS-derived isotope ratios: a comprehensive approach. Front. Microbiol. 9, 2342 (2018).
    Google Scholar 
    Svedén, J. B. et al. High cell-specific rates of nitrogen and carbon fixation by the cyanobacterium Aphanizomenon sp. at low temperatures in the Baltic Sea. FEMS Microbiol. Ecol. 91, fiv131 (2015).Meyer, N. R., Fortney, J. L. & Dekas, A. E. NanoSIMS sample preparation decreases isotope enrichment: magnitude, variability and implications for single-cell rates of microbial activity. Environ. Microbiol. 23, 81–89 (2021).
    Google Scholar 
    Woebken, D. et al. Revisiting N₂ fixation in Guerrero Negro intertidal microbial mats with a functional single-cell approach. ISME J. 9, 485–496 (2015).
    Google Scholar 
    Wickham, H. et al. Welcome to the tidyverse. J. Open Source Softw. 4, 1686 (2019).
    Google Scholar 
    De Mendiburu, F. agricolae: Statistical Procedures for Agricultural Research. https://CRAN.R-project.org/package=agricolae (2023).Jackman S. pscl: Classes and Methods for R Developed in the Political Science Computational Laboratory (University of Sydney, 2024).Zeileis, A., Kleiber, C. & Jackman, S. Regression Models for Count Data in R. J. Stat. Softw. 27, 1–25 (2008).Venables, W. N. & Ripley, B. D. Modern Applied Statistics with S-Plus 4th edn (Springer, 2002).Download referencesAcknowledgementsWe sincerely thank Christian Burmeister for nutrient analyses, Jenny Jeschek for DOC/DN analyses, as well as Christin Fechtel and the entire IOW monitoring team for mycoplankton sampling during the Baltic Sea monitoring cruises and at Heiligendamm. We further acknowledge the valuable support of Susanne Busch during phytoplankton identification, Heejin Jeon during IMS-1280 analyses, Sascha Plewe, as well as Karoline Schulz, Armin Springer, and Marcus Frank from the Medical Biology and Electron Microscopic Centre (Rostock University Medical Center, Rostock) during SEM imaging, Annett Grüttmüller during nanoSIMS analyses, and Quentin Devresse and Stanislav Jabinski during proofreading. A.F., C.D.L., and I.K. were funded by the German Research Foundation (DFG, Emmy Noether grant KL 3332/1-1 to IK). Sequencing was funded by the DFG (KL 3332/3-1 to IK) and conducted at the DFG Research Infrastructure NGS Competence Center (DcGC, project 407482635) as part of the Next Generation Sequencing Competence Network (DFG project 423957469). NGS library preparation, data production and analyses were carried out at the DcGC Dresden-concept Genome Center, core facility of the CMCB and Technology Platform of the Dresden University of Technology (TU Dresden). The NordSIMS facility is operated under a Swedish Research Council infrastructure grant 2021-00276. This is NordSIMS publication #814. The NanoSIMS at the Leibnitz-Institute for Baltic Sea Research, Warnemünde (IOW) was funded by the German Federal Ministry of Education and Research (BMBF, grant 03F0626A). The IOW long-term observations were financially supported by the IOW, the Federal Maritime and Hydrographic Agency (BSH), and the state and federal government.FundingOpen Access funding enabled and organized by Projekt DEAL.Author informationAuthors and AffiliationsDepartment of Biological Oceanography, Leibniz Institute for Baltic Sea Research, Rostock/Warnemuende, GermanyAnna Feuring, Connor D. Lawrence, Jessica Salcedo, Angela Vogts & Isabell KlawonnDepartment of Geosciences, Swedish Museum of Natural History, Stockholm, SwedenMartin J. WhitehouseCore Facility Bioinformatics, BOKU University, Vienna, AustriaLuca ZoccaratoInstitute of Computational Biology, BOKU University, Vienna, AustriaLuca ZoccaratoAuthorsAnna FeuringView author publicationsSearch author on:PubMed Google ScholarConnor D. LawrenceView author publicationsSearch author on:PubMed Google ScholarJessica SalcedoView author publicationsSearch author on:PubMed Google ScholarMartin J. WhitehouseView author publicationsSearch author on:PubMed Google ScholarAngela VogtsView author publicationsSearch author on:PubMed Google ScholarLuca ZoccaratoView author publicationsSearch author on:PubMed Google ScholarIsabell KlawonnView author publicationsSearch author on:PubMed Google ScholarContributionsI.K. acquired funding and designed research; A.F., C.D.L., J.S., M.J.W., A.V., and I.K. performed sampling and sample analyses; A.F., L.Z., and I.K. analyzed data; and A.F. and I.K. wrote the paper with input and approval from all coauthors.Corresponding authorCorrespondence to
    Isabell Klawonn.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Peer review

    Peer review information
    Nature Communications thanks the anonymous reviewer(s) 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 FileSupplementary Data 1Reporting SummaryTransparent Peer Review fileSource dataSource dataRights 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 articleFeuring, A., Lawrence, C.D., Salcedo, J. et al. Fungal parasites infecting N2-fixing cyanobacteria reshape carbon and N2 fixation and trophic transfer.
    Nat Commun (2026). https://doi.org/10.1038/s41467-025-67818-xDownload citationReceived: 11 June 2025Accepted: 09 December 2025Published: 02 January 2026DOI: https://doi.org/10.1038/s41467-025-67818-xShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    A One Health trial design to accelerate Lassa fever vaccines

    We propose an interdisciplinary framework to address the considerable challenges that are restricting the development of vaccines for zoonotic diseases with epidemic potential.

    Access through your institution

    Buy or subscribe

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

    Access options

    Access through your institution

    Access Nature and 54 other Nature Portfolio journals

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

    $32.99 / 30 days

    cancel any time

    Learn more

    Subscribe to this journal

    Receive 12 print issues and online access

    $259.00 per year
    only $21.58 per issue

    Learn more

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

    Additional access options:

    Log in

    Learn about institutional subscriptions

    Read our FAQs

    Contact customer support

    Fig. 1: The proposed OHTD framework.

    ReferencesWorld Health Organization. https://go.nature.com/3IUuBi2 (2024).Gouglas, D. et al. Lancet Glob. Health 6, e1386–e1396 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Warimwe, G. M. et al. Nat. Rev. Immunol. 21, 815–822 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Smith, D. R. M. et al. Nat. Med. 30, 3568–3577 (2024).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    World Health Organization. https://go.nature.com/4pRHM3Q (2017).Penfold, S. et al. PLoS ONE 18, e0283643 (2023).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Mariën, J. et al. J. Infect. Dis. https://doi.org/10.1093/infdis/jiaf308 (2025).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kenmoe, S. et al. PLoS Negl. Trop. Dis. 14, e0008589 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Andersen, K. G. et al. Cell 162, 738–750 (2015).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hoffmann, C. et al. Nat. Commun. 15, 9319 (2024).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Redding, D. W. et al. Nat. Commun. 12, 5759 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Latkin, C. A. et al. Addict. Behav. 73, 133–136 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Moore, S. M. et al. PLoS Negl. Trop. Dis. 19, e0012751 (2025).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Qian, Y. G., Jombart, T. & Edmunds, W. J. Vaccine X 14, 100321 (2023).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Capps, B. & Lederman, Z. J. Agric. Environ. Ethics 28, 1011–1032 (2015).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Download referencesAuthor informationAuthors and AffiliationsDepartment of Anthropology, The Pennsylvania State University, University Park, PA, USADavid Simons & Sagan FriantCenter for Infectious Disease Dynamics, The Pennsylvania State University, University Park, PA, USADavid Simons & Sagan FriantDepartment of Clinical Research, London School of Hygiene and Tropical Medicine, London, UKDeborah Watson-Jones & Daniela MannoMwanza Intervention Trials Unit, National Institute for Medical Research, Mwanza, TanzaniaDeborah Watson-JonesAuthorsDavid SimonsView author publicationsSearch author on:PubMed Google ScholarDeborah Watson-JonesView author publicationsSearch author on:PubMed Google ScholarDaniela MannoView author publicationsSearch author on:PubMed Google ScholarSagan FriantView author publicationsSearch author on:PubMed Google ScholarCorresponding authorCorrespondence to
    David Simons.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Rights and permissionsReprints and permissionsAbout this articleCite this articleSimons, D., Watson-Jones, D., Manno, D. et al. A One Health trial design to accelerate Lassa fever vaccines.
    Nat Med (2026). https://doi.org/10.1038/s41591-025-04018-6Download citationPublished: 02 January 2026Version of record: 02 January 2026DOI: https://doi.org/10.1038/s41591-025-04018-6Share 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

    The use of multi-sensor drone data for the development and validation of methods to track and characterize marine animals

    Abstract

    Low cost, unmodified, commercially available drones can provide an effective platform for the study and characterization of marine megafauna. We present methods which utilize video and flight data to allow for both the continuous tracking of animals and the determination of animal lengths across a range of flight parameters. We also provide a thorough estimation of error in animal position and length measurements while at the same time introducing methods to correct for errors in reported aircraft altitude and heading. Methods are validated using both ground-based markers and tracking data from free swimming white sharks which includes the simultaneous tracking of individual sharks by two drones as the aircraft undergo changes in altitude, gimbal angle, heading and position. The resultant tracks are seen to be highly congruent (mean distance between measured positions: 4.3 m (95% CI 0 to 10)) and length measurements demonstrate a high level of precision (95% CI −8 to 8%) with accuracy confirmed using ground-based markers (mean error: 0.3% (95% CI −4.8 to 4.8%)). Results demonstrate the effectiveness of these methods across a range of flight conditions encountered in the field. The methods introduced allow flexibility in data capture while still providing accurate information, with the potential to both expand the use of and enhance the value of drone-based data for the quantification of animal behaviors and characteristics.

    Data availability

    All processed data reported in this study as well as aligned flight and image sensor data for reported shark tracks and ground targets, along with the code used to generate figures and conduct analyses will be made available via Code Ocean upon publication. Raw video and flight log data are not publicly archived at this time. However, reasonable requests for access to specific subsets of the raw data that support the findings of this study will be considered by the corresponding author.
    ReferencesButcher, P. A. et al. The drone revolution of shark science: A review. Drones 5, 8 (2021).
    Google Scholar 
    Rex, P. T., May, J. H. III, Pierce, E. K. & Lowe, C. G. Patterns of overlapping habitat use of juvenile white shark and human recreational water users along Southern California beaches. PLoS One. 18, e0286575 (2023).
    Google Scholar 
    Colefax, A. P., Kelaher, B. P., Pagendam, D. E. & Butcher, P. A. Assessing white shark (Carcharodon carcharias) behavior along coastal beaches for Conservation-Focused shark mitigation. Front. Mar. Sci. 7, 519539 (2020).
    Google Scholar 
    Raoult, V., Tosetto, L. & Williamson, J. E. Drone-Based High-Resolution tracking of aquatic vertebrates. Drones 2, 37 (2018).
    Google Scholar 
    Tanaka, K. R. et al. North Pacific warming shifts the juvenile range of a marine apex predator. Sci. Rep. 11, 3373 (2021).
    Google Scholar 
    Raoult, V. et al. Operational protocols for the use of drones in marine animal research. Drones 4, 64 (2020).
    Google Scholar 
    Bierlich, K. C. et al. Comparing uncertainty associated with 1-, 2-, and 3D aerial Photogrammetry-Based body condition measurements of Baleen whales. Front. Mar. Sci. 8 https://doi.org/10.3389/fmars.2021.749943 (2021).Bierlich, K. C. et al. Bayesian approach for predicting photogrammetric uncertainty in morphometric measurements derived from drones. Mar. Ecol. Prog Ser. 673, 193–210 (2021).
    Google Scholar 
    Jech, J. M. et al. Measurements of juvenile Atlantic Bluefin tuna (Thunnus thynnus) size using an unmanned aerial system. J. Unmanned Veh. Syst. 8, 140–160 (2020).
    Google Scholar 
    Burnett, J. D. et al. Estimating morphometric attributes of Baleen whales with photogrammetry from small uass: A case study with blue and Gray whales. Mar. Mamm. Sci. 35, 108–139 (2019).
    Google Scholar 
    Soledade Lemos, L., Burnett, J. D., Chandler, T. E., Sumich, J. L. & Torres, L. G. Intra- and inter‐annual variation in Gray Whale body condition on a foraging ground. Ecosphere 11 https://doi.org/10.1002/ecs2.3094 (2020).Vivier, F. et al. Quantifying the age structure of free-ranging delphinid populations: testing the accuracy of unoccupied aerial system photogrammetry. Ecol. Evol. 13, e10082 (2023).
    Google Scholar 
    Dawson, S. M., Bowman, M. H., Leunissen, E. & Sirguey, P. Inexpensive aerial photogrammetry for studies of whales and large marine animals. Front. Mar. Sci. 4 https://doi.org/10.3389/fmars.2017.00366 (2017).LightWare SF11 Laser Altimeter Product Manual. https://lightwarelidar.com/wp-content/uploads/2025/07/SF11-Laser-Altimeter-Manual-Rev-10.pdf (2021).Randall, J. E. Size of the great white shark (carcharodon). Science 181, 169–170 (1973).
    Google Scholar 
    Webster, D. K. Myth and Maneater: the Story of the Shark (Norton, 1963).Irschick, D. J. Measuring performance in nature: implications for studies of fitness within populations. Integr. Comp. Biol. 43, 396–407 (2003).
    Google Scholar 
    Arnold, S. J. Morphology, performance and Fitness1. Integr. Comp. Biol. 23, 347–361 (2015).
    Google Scholar 
    Skomal, G. B., Braun, C. D., Chisholm, J. H. & Thorrold, S. R. Movements of the white shark Carcharodon carcharias in the North Atlantic ocean. Mar. Ecol. Prog Ser. 580, 1–16 (2017).
    Google Scholar 
    Krebs, J. R. & Stephens, D. W. Foraging Theory (Princeton University Press, 2019).Support for Phantom 4. In: DJI [Internet]. [cited 18 Apr 2024]. https://www.dji.com/support/product/phantom-4Watanabe, Y. Y. & Takahashi, A. Linking animal-borne video to accelerometers reveals prey capture variability. Proc. Natl. Acad. Sci. U S A. 110, 2199–2204 (2013).
    Google Scholar 
    Watanabe, Y. Y., Payne, N. L., Semmens, J. M., Fox, A. & Huveneers, C. Hunting behaviour of white sharks recorded by animal-borne accelerometers and cameras. Mar. Ecol. Prog Ser. 621, 221–227 (2019).
    Google Scholar 
    Born, M. & Wolf, E. Principles of Optics: 60th Anniversary Edition (Cambridge University Press, 2019).Comaniciu, D., Ramesh, V. & Meer, P. Real-time tracking of non-rigid objects using mean shift. In Proceedings IEEE Conference on Computer Vision and Pattern Recognition CVPR 2000 (Cat NoPR00662). IEEE Comput. Soc. https://doi.org/10.1109/cvpr.2000.854761 (2002).Download referencesAcknowledgementsThe authors would like to thank Mark McCormick, David Sexton, Taylor Lamme, David Mizrahi, Malcolm Stewart, Josh Stewart, Geoff Reuland, John Reuland, and Jackson R. Davis for their assistance with field experiments and piloting. They also wish to thank Max Kullberg and scientists at OCEARCH who provided valuable insight in the preparation of this manuscript. The authors would like to further acknowledge support from the Kay Fellowship through the ‘Iolani School. Finally, they would like to thank the Cape Cod Ocean Community and the Dorr Foundation for their support.Author informationAuthors and AffiliationsMoosh Systems, LLC, Wellfleet, MA, USAKristian J. Sexton, Elise F. Danko, Austin Boesch, Nicholas Cheung & Robert W. HoltThe Cooper Union for the Advancement of Science and Art, New York, NY, USAElise F. DankoBiology Department, Stanford University, Stanford, CA, USAAlexandra E. DiGiacomoAuthorsKristian J. SextonView author publicationsSearch author on:PubMed Google ScholarElise F. DankoView author publicationsSearch author on:PubMed Google ScholarAustin BoeschView author publicationsSearch author on:PubMed Google ScholarNicholas CheungView author publicationsSearch author on:PubMed Google ScholarAlexandra E. DiGiacomoView author publicationsSearch author on:PubMed Google ScholarRobert W. HoltView author publicationsSearch author on:PubMed Google ScholarContributionsKJS planned the study and developed the computational methods used in data analysis. KJS and EFD conceived, designed and carried out experiments. RH and EFD provided conceptual input throughout planning, testing and data analysis. KJS, EFD and RH performed data analysis. KJS wrote the initial draft of the manuscript. EFD, RH, NC, AED, and AB provided feedback, editing and numerous other contributions to the further development of the manuscript.Corresponding authorCorrespondence to
    Kristian J. Sexton.Ethics declarations

    Competing interests
    KJS declares the following competing interests: KJS has filed a patent entitled ‘Method to correct for reported drone altitude errors over duration of a flight based on determination of true altitude error from at least one point within the flight’ with the US Patent and Trademark Office, which covers the landing altitude correction method described in this paper. KJS has filed a patent entitled “System and Method for Correcting Drone Heading and Data Misalignment Using Path-Based and Correlation-Based Optimization” with the US Patent and Trademark Office, which covers the flight and image data alignment method described in this paper. KJS has also filed a provisional patent entitled ‘Method for correcting reported UAS heading errors using single or multiple offset corrections’ with the US Patent and Trademark Office, which covers the heading error correction methods described in this paper. The remaining 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 articleSexton, K.J., Danko, E.F., Boesch, A. et al. The use of multi-sensor drone data for the development and validation of methods to track and characterize marine animals.
    Sci Rep (2026). https://doi.org/10.1038/s41598-025-31975-2Download citationReceived: 06 February 2025Accepted: 06 December 2025Published: 02 January 2026DOI: https://doi.org/10.1038/s41598-025-31975-2Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative
    KeywordsAnimal trackingDrone-based measurementsDrone accuracyMarine megafaunaQuantitative imagingWhite shark More

  • in

    Blue carbon sequestration dominated by dissolved organic carbon pathways for kelp forests and eelgrass meadows in Nova Scotia, Canada

    AbstractExisting estimates of carbon sequestration by coastal vegetated ecosystems (e.g., kelp forests, seagrass meadows) remain highly uncertain because they utilize few measurements with high uncertainty or focus on individual stocks or fluxes. Here, we combine empirical data with modeling to generate detailed carbon budgets for kelp forests and eelgrass meadows in Nova Scotia, Canada. Our budgeting showed that the release and export of dissolved organic carbon accounts for substantially more carbon sequestered by these habitats than has been estimated globally (98.5 ± 64.6% for kelp and 84 ± 36.0% for eelgrass), suggesting this carbon pathway has been undervalued. Further, kelp forests are estimated to sequester more carbon than eelgrass by ~1.3 orders of magnitude (27 ± 19% and 12.1 ± 6.3% of annual net primary production for kelp and eelgrass) mainly due to their large habitat area, highlighting their significant but largely overlooked role in ocean carbon sequestration relative to other blue carbon ecosystems.

    Data availability

    All data are available as supplementary material and at: https://github.com/kirakrumhansl/Carbon-Budget-Code.
    Code availability

    All code and associated raw data are available at: https://github.com/kirakrumhansl/Carbon-Budget-Code.
    ReferencesNellemann, C. et al. (eds) Blue Carbon: A Rapid Response Assessment (United Nations Environment Programme, GRID-Arendal, 2009).Duarte, C. M. & Krause-Jensen, D. Export from seagrass meadows contributes to marine carbon sequestration. Front. Mar. Sci. 4, 13 (2017).Krause-Jensen, D. & Duarte, C. M. Substantial role of macroalgae in marine carbon sequestration. Nat. Geosci. 9, 737–742 (2016).
    Google Scholar 
    Trevathan-Tackett, S. M. et al. Effects of small-scale, shading-induced seagrass loss on blue carbon storage: Implications for management of degraded seagrass ecosystems. J. Appl. Ecol. 55, 1351–1359 (2018).
    Google Scholar 
    Blain, C. O., Hansen, S. C. & Shears, N. T. Coastal darkening substantially limits the contribution of kelp to coastal carbon cycles. Glob. Change Biol. 27, 5547–5563 (2021).
    Google Scholar 
    Drever, R. C. et al. Natural climate solutions for Canada. Sci. Adv. 7, eabd6034 (2021).
    Google Scholar 
    Pessarrodona, A. et al. Carbon sequestration and climate change mitigation using macroalgae: a state of knowledge review. Biol. Rev. Camb. Philos. Soc. 98, 1945–1971 (2023).
    Google Scholar 
    Macreadie, P. I. et al. The future of Blue Carbon science. Nat. Commun. 10, 3998 (2019).
    Google Scholar 
    Filbee-Dexter, K. et al. Carbon export from seaweed forests to deep ocean sinks. Nat. Geosci. 17, 552–559 (2024).
    Google Scholar 
    Serrano, O. et al. Australian vegetated coastal ecosystems as global hotspots for climate change mitigation. Nat. Commun. 10, 4313 (2019).
    Google Scholar 
    Reynolds, L. K. et al. Ecosystem services returned through seagrass restoration. Restor. Ecol. 24, 583–588 (2016).
    Google Scholar 
    Huang, Y. H. et al. Carbon budgets of multispecies seagrass beds at Dongsha Island in the South China Sea. Mar. Environ. Res. 106, 92–102 (2015).
    Google Scholar 
    Frigstad, H. et al. Blue Carbon – Climate Adaptation, $CO_2$ Uptake and Sequestration of Carbon in Nordic Blue Forests (NIVA Report 7448-2020, Norwegian Institute for Water Research, 2020).Watanabe, K. et al. Macroalgal metabolism and lateral carbon flows can create significant carbon sinks. Biogeosciences 17, 2425–2440 (2020).
    Google Scholar 
    Macreadie, P. I. et al. Blue carbon as a natural climate solution. Nat. Rev. Earth Environ. 2, 826–839 (2021).
    Google Scholar 
    O’Brien, J. M., Wong, M. C. & Stanley, R. R. E. Fine-scale ensemble species distribution modeling of eelgrass (Zostera marina) to inform nearshore conservation planning and habitat management. Front. Mar. Sci. 9, 988858 (2022).Krumhansl, K. A. et al. Pathways of blue carbon export from kelp and seagrass beds along the Atlantic coast of Nova Scotia. Sci. Adv. 11, eadw1952 (2025).
    Google Scholar 
    Fourqurean, J. W. et al. Seagrass ecosystems as a globally significant carbon stock. Nat. Geosci. 5, 505–509 (2012).
    Google Scholar 
    Röhr, M. E. et al. Blue carbon storage capacity of temperate Eelgrass (Zostera marina) Meadows. Glob. Biogeochem. Cycles 32, 1457–1475 (2018).
    Google Scholar 
    Novak, A. B. et al. Factors influencing carbon stocks and accumulation rates in eelgrass meadows across New England, USA. Estuaries Coasts 43, 2076–2091 (2020).
    Google Scholar 
    Song, H. et al. Interannual variability in phytoplankton blooms and plankton productivity over the Nova Scotian Shelf and in the Gulf of Maine. Mar. Ecol. Prog. Ser. 426, 105–118 (2011).
    Google Scholar 
    Krumhansl, K. A. & Scheibling, R. E. Detrital production in Nova Scotian kelp beds: patterns and processes. Mar. Ecol. Prog. Ser. 421, 67–82 (2011).
    Google Scholar 
    Nienhuis, P. H. & Groenendijk, A. M. Consumption of eelgrass (Zostera marina) by birds and invertebrates: an annual budget. Mar. Ecol. Prog. Ser. 29, 29–35 (1986).
    Google Scholar 
    Krumhansl, K. A. & Scheibling, R. E. Production and fate of kelp detritus. Mar. Ecol. Prog. Ser. 467, 281–302 (2012).
    Google Scholar 
    Filbee-Dexter, K. & Scheibling, R. E. Hurricane-mediated defoliation of kelp beds and pulsed delivery of kelp detritus to offshore sedimentary habitats. Mar. Ecol. Prog. Ser. 455, 51–64 (2012).
    Google Scholar 
    Hansell, D. A. & Carlson, C. A. (eds) Biogeochemistry of Marine Dissolved Organic Matter (Academic Press, 2002).Pedersen, M. F. et al. Carbon sequestration potential increased by incomplete anaerobic decomposition of kelp detritus. Mar. Ecol. Prog. Ser. 660, 53–67 (2021).
    Google Scholar 
    Silverberg, N. et al. Remineralization of organic carbon in eastern Canadian continental margin sediments. Deep-Sea Res. II 47, 699–731 (2000).
    Google Scholar 
    Ortega, A. et al. Connected macroalgal–sediment systems: blue carbon and food webs in the deep coastal ocean. Ecol. Monogr. 89, e01366 (2019).Kuwae, T. & Hori, M. (eds) Blue Carbon in Shallow Coastal Ecosystems: Carbon Dynamics, Policy, and Implementation (Springer Nature, 2019).Trevathan-Tackett, S. M. et al. Long-term decomposition captures key steps in microbial breakdown of seagrass litter. Sci. Total Environ. 705, 135806 (2020).
    Google Scholar 
    Weatherall, E. J. et al. Quantifying the dispersal potential of seagrass vegetative fragments: a comparison of multiple subtropical species. Estuar., Coast. Shelf Sci. 169, 207–215 (2016).
    Google Scholar 
    Chang, Y. E. et al. Lateral carbon export of macroalgae and seagrass from diverse habitats contributes particulate organic carbon in the deep sea of the Northern South China Sea. Mar. Pollut. Bull. 206, 116672 (2024).
    Google Scholar 
    Paine, E. R. et al. Rate and fate of dissolved organic carbon release by seaweeds: a missing link in the coastal ocean carbon cycle. J. Phycol. 57, 1375–1391 (2021).
    Google Scholar 
    Pellikaan, G. C. & Nienhuis, P. H. Nutrient uptake and release during growth and decomposition of eelgrass, Zostera marina L., and its effects on the nutrient dynamics of Lake Grevelingen. Aquat. Bot. 30, 189–214 (1988).
    Google Scholar 
    Gao, Y. et al. Dissolved organic carbon from cultured kelp Saccharina japonica: production, bioavailability, and bacterial degradation rates. Aquac. Environ. Interact. 13, 101–110 (2021).
    Google Scholar 
    Egea, L. G. et al. Marine heatwaves and disease alter community metabolism and DOC fluxes on a widespread habitat-forming seagrass species (Zostera marina). Sci. Total Environ. 957, 177820 (2024).
    Google Scholar 
    Pellikaan, G. C. Laboratory Experiments on Eelgrass (Zostera marina L.) Decomposition. Neth. J. Sea Res. 18, 360–383 (1984).
    Google Scholar 
    Perkins, A. K. et al. Production of dissolved carbon and alkalinity during macroalgal wrack degradation on beaches: a mesocosm experiment with implications for blue carbon. Biogeochemistry 160, 159–175 (2022).
    Google Scholar 
    Godshalk, G. L. & Wetzel, R. G. Decomposition of aquatic angiosperms. III. Zostera marina and a conceptual model of decomposition. Aquat. Bot. 5, 329–354 (1978).
    Google Scholar 
    Liu, X. et al. Simulating water residence time in the coastal ocean: a global perspective. Geophys. Res. Lett. 46, 13910–13919 (2019).
    Google Scholar 
    Hill, R. et al. Can macroalgae contribute to blue carbon? An Australian perspective. Limnol. Oceanogr. 60, 1689–1706 (2015).
    Google Scholar 
    Filbee-Dexter, K. & Wernberg, T. Substantial blue carbon in overlooked Australian kelp forests. Sci. Rep. 10, 12341 (2020).
    Google Scholar 
    Franco, J. N. et al. Potential blue carbon in the fringe of Southern European Kelp forests. Sci. Rep. 15, 29573 (2025).
    Google Scholar 
    van der Mheen, M. et al. Substantial kelp detritus exported beyond the continental shelf by dense shelf water transport. Sci. Rep. 14, 839 (2024).
    Google Scholar 
    Johannessen, S. C. How to quantify blue carbon sequestration rates in seagrass meadow sediment: geochemical method and troubleshooting. Carbon Footprints 2, 5 (2023).Krumhansl, K. A. et al. Loss, resilience and recovery of kelp forests in a region of rapid ocean warming. Ann. Bot. 133, 73–92 (2024).
    Google Scholar 
    Andraos, J. On the propagation of statistical errors for a function of several variables. J. Chem. Educ. 73, 150–154 (1996).
    Google Scholar 
    Mann, K. H. Ecological Energetics of the seaweed zone in a marine bay on the Atlantic Coast of Canada. I. Zonation and biomass of seaweeds. Mar. Biol. 12, 1–10 (1972).
    Google Scholar 
    Wong, M. C. & M. Dowd. The role of short-term temperature variability and light in shaping the phenology and characteristics of seagrass beds. Ecosphere 14, e4461 (2023).Postlethwaite, V. R. et al. Low blue carbon storage in eelgrass (Zostera marina) meadows on the Pacific Coast of Canada. PLoS One 13, e0198348 (2018).
    Google Scholar 
    Christensen, M. S. Estimating blue carbon storage capacity of Canada’s eelgrass beds. MSc thesis, Univ. British Columbia (2023).Pessarrodona, A. et al. A global dataset of seaweed net primary productivity. Sci. Data 9, 484 (2022).
    Google Scholar 
    Hatcher, B. G. & Chapman, A. R. O. & Mann, K.H. An Annual Carbon Budget for the Kelp Laminaria longicruris. Mar. Biol. 44, 85–96 (1977).
    Google Scholar 
    Gaeckle, J. L. & Short, F. T. A plastochrone method for measuring leaf growth in Eelgrass, Zostera marina. Bull. Mar. Sci. 71, 1237–1246 (2002).
    Google Scholar 
    Short, F. T. & Duarte, C. M. Methods for the measurement of seagrass growth and production. In Global Seagrass Research Methods (eds Short, F. T. & Coles, R. G.) 155–182 (Elsevier Science, 2001).Thomson, J. A., Vercaemer, B. & Wong, M. C. Non-destructive biomass estimation for eelgrass (Zostera marina): Allometric and percent cover–biomass relationships vary with environmental conditions. Aquat. Bot. 198, 103833 (2025).Chapman, A. R. O. Reproduction, recruitment, and mortality in two species of Laminaria in southwest Nova Scotia. J. Exp. Mar. Biol. Ecol. 78, 99–109 (1984).
    Google Scholar 
    Hebert, D. et al. Physical Oceanographic Conditions on the Scotian Shelf and in the Gulf of Maine during 2023. Can. Tech. Rep. Hydrogr. Ocean Sci. 380, 71 (2024).Grebe, G. S. et al. The effect of distal-end trimming onSaccharina latissimamorphology, composition, and productivity. J. World Aquac. Soc. 52, 1081–1098 (2021).
    Google Scholar 
    Ronowicz, M., Kukliński, P. & Włodarska-Kowalczuk, M. Morphological variation of kelps (Alaria esculenta, cf. Laminaria digitata, and Saccharina latissima) in an Arctic glacial fjord. Estuar. Coast. Shelf Sci. 268, 107801 (2022).Romero, J. et al. The detritic compartment in a posidonia oceanica meadow: litter features, decomposition rates, and mineral stocks. Mar. Ecol. 13, 69–83 (2008).
    Google Scholar 
    Reed, D. C. et al. Patterns and controls of reef-scale production of dissolved organic carbon by giant kelp Macrocystis pyrifera. Limnol. Oceanogr. 60, 1996–2008 (2015).
    Google Scholar 
    Weigel, B. L. & Pfister, C. A. The dynamics and stoichiometry of dissolved organic carbon release by kelp. Ecology 102, e03221 (2021).
    Google Scholar 
    Jimenez-Ramos, R. et al. Carbon metabolism and bioavailability of dissolved organic carbon (DOC) fluxes in seagrass communities are altered under the presence of the tropical invasive alga Halimeda incrassata. Sci. Total Environ. 839, 156325 (2022).
    Google Scholar 
    Jones, E., Zemlyak, F. & Stewart, P. Operating manual for the Bedford Institute of Oceanography Automated Dissolved Oxygen Titration System (Can. Tech. Rep. Hydrogr. Ocean Sci. 138, 1992).Lee, K., Park, S. & Kim, Y. Effects of irradiance, temperature, and nutrients on growth dynamics of seagrasses: a review. J. Exp. Mar. Biol. Ecol. 350, 144–175 (2007).
    Google Scholar 
    Wong, M. C. & Dowd, M. Eelgrass (Zostera marina) trait variation across varying temperature–light regimes. Estuaries Coasts 48, 15 (2025).Wickham, H. ggplot2: Elegant Graphics for Data Analysis (Springer-Verlag New York, 2016).Pante, E., Simon-Bouhet, B._marmap: import, plot and analyze bathymetric and topographic data. CRAN https://doi.org/10.32614/CRAN.package.marmap (2025).Download referencesAcknowledgementsWe acknowledge Katie Thistle, Shawn Roach, Thomas Baker, Cody Brooks, Danielle Davenport, Claudio DiBacco, Katherine Lee, Jordan Thomson, Benedikte Vercaemer, Chris Corriveau, and Nick Jeffery for their support in the field. We also thank Wendy Gentleman for early input and feedback on the work. KK discloses support for this work from a Fisheries and Oceans Canada Competitive Science Research Fund grant to KK and KA-S.Author informationAuthors and AffiliationsBedford Institute of Oceanography, Fisheries and Oceans Canada, Dartmouth, NS, CanadaKira A. Krumhansl, Melisa C. Wong, Manon M. M. Picard, Meredith Fraser, Carrie-Ellen Gabriel, Yongsheng Wu & Kumiko Azetsu-ScottAuthorsKira A. KrumhanslView author publicationsSearch author on:PubMed Google ScholarMelisa C. WongView author publicationsSearch author on:PubMed Google ScholarManon M. M. PicardView author publicationsSearch author on:PubMed Google ScholarMeredith FraserView author publicationsSearch author on:PubMed Google ScholarCarrie-Ellen GabrielView author publicationsSearch author on:PubMed Google ScholarYongsheng WuView author publicationsSearch author on:PubMed Google ScholarKumiko Azetsu-ScottView author publicationsSearch author on:PubMed Google ScholarContributionsK.K. was involved in conceptualization, data curation, formal analysis, funding acquisition, investigation, methodology, project administration, resources, software, supervision, visualization, writing—original draft, writing—reviewing & editing. M.W. was involved in conceptualization, data curation, formal analysis, funding acquisition, investigation, methodology, project administration, resources, software, supervision, writing—original draft, writing—reviewing & editing. M.P. was involved in data curation, formal analysis, investigation, methodology, software, writing—original draft, writing—reviewing & editing. M.F. was involved in data curation, formal analysis, investigation, software, writing—original draft, writing—reviewing & editing. C.–E.G. Gabriel was involved in investigation, methodology, writing—reviewing & editing. Y.W. was involved in Methodology and writing—reviewing & editing. K.A.-S. was involved in conceptualization, methodology, project administration, resources, supervision, writing—reviewing & editing.Corresponding authorCorrespondence to
    Kira A. Krumhansl.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Peer review

    Peer review information
    Communications Earth and Environment thanks Richard Zimmerman and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Primary Handling Editors: Nezha Mejjad and Alice Drinkwater. A peer review file is available.

    Additional informationPublisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary informationSupplemental MaterialTransparent 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 articleKrumhansl, K.A., Wong, M.C., Picard, M.M.M. et al. Blue carbon sequestration dominated by dissolved organic carbon pathways for kelp forests and eelgrass meadows in Nova Scotia, Canada.
    Commun Earth Environ (2026). https://doi.org/10.1038/s43247-025-03122-2Download citationReceived: 09 May 2025Accepted: 10 December 2025Published: 02 January 2026DOI: https://doi.org/10.1038/s43247-025-03122-2Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Europeans like wolves, lynx and bears and oppose hunting them but do not want more of these animals

    A Europe‑wide survey of 10,000 respondents shows strong backing for the recovery of wolves, lynx and bears, as well as opposition to hunting and culling, but also shows that a majority opposes any further population growth. Many respondents remain neutral, which indicates that large carnivore recovery is far less polarizing than commonly portrayed.

    Access through your institution

    Buy or subscribe

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

    Access options

    Access through your institution

    Access Nature and 54 other Nature Portfolio journals

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

    $32.99 / 30 days

    cancel any time

    Learn more

    Subscribe to this journal

    Receive 12 digital issues and online access to articles

    $119.00 per year
    only $9.92 per issue

    Learn more

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

    Additional access options:

    Log in

    Learn about institutional subscriptions

    Read our FAQs

    Contact customer support

    Fig. 1: Public opinion in the EU concerning the conservation and management of large carnivores.

    ReferencesChapron, G. et al. Recovery of large carnivores in Europe’s modern human-dominated landscapes. Science 346, 1517–1519 (2014). A review that describes the European continent-wide recovery of large carnivores.Article 
    CAS 
    PubMed 

    Google Scholar 
    Di Bernardi, C. et al. Continuing recovery of wolves in Europe. PLoS Sustain. Transform. 4, e0000158 (2025). A review that presents an update on the continuing recovery of wolves in Europe.Article 

    Google Scholar 
    Fleurke, F. & Trouwborst, A. On an anti-wolf mission, commission ignores science and law. Eur. Law Blog https://doi.org/10.21428/9885764c.a25018f4 (2025). A blog post that discusses controversies about wolf downlisting in the Habitats Directive.Article 

    Google Scholar 
    Swedish Government. We prioritise reducing the wolf population (in Swedish). regeringen.se https://www.regeringen.se/debattartiklar/2025/04/vi-prioriterar-en-minskad-vargstam/ (2025). An opinion piece by Swedish ministers that announces their intention to halve the wolf population.Liddell, T. M. & Kruschke, J. K. Analyzing ordinal data with metric models: what could possibly go wrong? J. Exp. Soc. Psychol. 79, 328–348 (2018). A paper that explains how type I and II errors can arise when treating ordinal data as metric.Article 

    Google Scholar 
    Download referencesAdditional informationPublisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.This is a summary of: Chapron, G. et al. Europeans support large carnivore recovery while opposing both further population growth and hunting. Nat. Ecol. Evol. https://doi.org/10.1038/s41559-025-02914-1 (2025).Rights and permissionsReprints and permissionsAbout this articleCite this article Europeans like wolves, lynx and bears and oppose hunting them but do not want more of these animals.
    Nat Ecol Evol (2026). https://doi.org/10.1038/s41559-025-02915-0Download citationPublished: 02 January 2026Version of record: 02 January 2026DOI: https://doi.org/10.1038/s41559-025-02915-0Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Examining anomalous summer carbonyl sulfide emissions in a boreal forest after thinning

    AbstractCarbonyl sulfide (COS) is gaining interest as a proxy for gross primary productivity (GPP). Thinning of the Hyytiälä (Finland) forest in the winter of 2019–2020 altered the response of COS fluxes to environmental conditions in the summer of 2021. For the first time, extended periods of ecosystem-scale COS emissions were observed in a boreal forest. The warm and dry conditions in the summer of 2021 reduced the COS uptake by the canopy and elevated soil abiotic COS production. However, the reduction in canopy uptake and the increase in soil production do not fully explain the observed ecosystem-level emissions. The analysis suggests an unidentified, homogeneously distributed COS source in the eddy covariance footprint area, potentially from the photodegradation of forest floor litter and cutting residue from thinning. Such a source in a boreal forest stand warrants further source apportionment studies to effectively use COS as a proxy for GPP.

    IntroductionCarbonyl sulfide (COS) is an atmospheric tracer for ecosystem carbon and water fluxes because the diffusive pathway for COS is similar to that for carbon dioxide (CO2)1,2,3,4,5. COS, along with CO2, is taken up through the stomata to the chloroplast surface, where COS is irreversibly hydrolysed by carbonic anhydrase (CA)6,7,8. While CO2 is emitted back to the atmosphere during plant respiration, COS emission by leaves is not observed under ambient conditions7,9,10. The key assumption behind using COS as a proxy for ecosystem-level gross primary production (GPP) is that photosynthesising plants are a major terrestrial sink of COS.In the boreal forest ecosystem, canopy uptake is typically the largest COS flux, even at night, followed by soil uptake11. The incomplete closure of stomata and the light-independent nature of CA lead to the nighttime uptake of COS12,13,14, which correlates with stomatal conductance11,15. During the day, stomatal conductance linearly increases with light intensity in a limited range until it plateaus, and COS uptake could be used to derive canopy and stomatal conductance, given that soil fluxes could be distinguished from separate measurements13,16,17.In order to accurately use COS as a GPP proxy, all the non-stomatal sinks and sources have to be quantified. The low seasonal variation of COS soil fluxes in the boreal forest makes it easier to characterise the canopy uptake and account for ecosystem budgets and GPP estimates18. Soil COS chamber measurements at Hyytiälä forest station have shown an average flux of ~ −3 pmol m−2 s−1 during the growing season, with weak seasonal changes18. Soil uptake is linked to the presence of CA in microbial communities19,20, necessitating optimal soil temperature and moisture conditions for the maximal absorption of COS by soil18,21,22,23,24. The contribution of soil to the global COS sink is estimated to be 26–33%1,25 but is lower in boreal forests (10–20%)5,18.Wetlands and anoxic soils are considered terrestrial sources of COS20,26,27,28, while oxic soils are generally inferred to be COS sinks from measurements18,19,29 and modelling studies24,30,31. However, exposure to high temperatures (above 15 °C) and high insolation in dry conditions can result in COS emissions from oxic soils14,21,32,33,34. Soil COS emissions are primarily linked to the oxic and abiotic degradation of organic matter within the soil34, and in the case of anoxic soils, redox reactions are important21,26. Furthermore, emission of COS due to exposure to UV radiation was shown to vary with the amount of soil organic matter34,35. Agriculturally managed temperate mountain grasslands turned into a net source of COS shortly after grass cutting, and the soil contribution could not explain the net ecosystem level emissions36. Commane et al.37 observed unexpected net COS emission over a mid-latitude forest site during the warmest summer weeks.To better understand and quantify the source and sink processes in various ecosystems, and thereby improve the accuracy of GPP estimates, COS measurements in a variety of ecosystems at longer temporal scales are needed1,38,39,40. The bottom-up global estimate of COS is not closed, suggesting an unaccounted source of COS in the tropics41. Inverse modelling studies have also shown a missing source in the tropics and a missing sink in the high latitudes42,43,44. Parameterising seasonal and interannual variation of the boreal forest COS uptake in the Simple Biosphere Model Version 4 (SiB4) was able to match the missing COS sink to a certain limit in the high latitude, but the regional variation remains unknown5. A recent measurement has found boreal wetlands to be a sink for COS45, which was previously reported as a source4,26,27, underscoring the complexity of ecosystem processes in the global COS budget.COS fluxes at the ecosystem level can be studied using the eddy covariance (EC) method at high temporal resolution. Several EC flux measurements of COS and attempts to estimate GPP from them have been reported in the literature5,13,37,46,47. The most extensive EC measurements of COS flux from a boreal forest were conducted at the Hyytiälä forest station, spanning over 32 months across five years from 2013 to 20175. This multi-year dataset showed that monthly median COS uptake varied by up to 26%. The main controlling bioclimatic factors of COS fluxes at the ecosystem level were photosynthetically active radiation (PAR), vapour pressure deficit (VPD), air temperature (Ta), and leaf area index (LAI).In this study, we measured and analysed three more years (2020–2022) of COS flux measurements at the Hyytiälä forest station following a thinning conducted in 2019–2020 that removed ~ 40% basal area. Building on the well-studied COS dynamics of Hyytiälä forest station, we leveraged this unique ecosystem manipulation in the boreal forest to better understand the role of different ecosystem components in COS fluxes. Thinning altered the microclimate of the forest floor and created a large amount of cutting residue that was left on the site48. The post-thinning COS fluxes exhibited anomalous behaviour, and in the summer of 2021, we observed continuous and long-lasting ecosystem-level COS emissions from the boreal forest. Here, we quantify the emission of COS, identify the various potential sources of COS and understand the drivers that controlled the anomalous COS emission in the boreal forest ecosystem.Results and discussionsAnomalous COS emissions and meteorological conditionsThinning reduced forest canopy COS uptake, determined from the net ecosystem COS fluxes (FCOS) measured by EC. FCOS was ~19 % lower in 2020 (−19.6 ± 13.4 pmol m−2 s−1; mean ± standard deviation) and nearly 24 % lower in 2022 (−18.4 ± 13.1 pmol m−2 s−1) compared to the pre-thinning five-year (2013–17) average of −24.2 ± 13.3 pmol m−2 s−1. However, FCOS was about 75 % lower in 2021 (−6.0 ± 14.5 pmol m−2 s−1) due to anomalous COS emissions that were observed from June to mid-August and peaked in June and July (Fig. 1). Consequently, this study focuses on these months to understand and isolate the contributing conditions. The seasonal variation (FCOS) remained consistent in the post-thinning years, with maximum uptake in summer months and minimal uptake in winter months (Supplementary Fig. 1).Fig. 1: Diel variation of environmental variables and FCOS over the months.Mean diel variation of photosynthetically active radiation (PAR), air temperature average over heights of 16.8 m and 33.6 m (Ta), vapour pressure deficit (VPD), soil temperature (Ts) of horizon B1 (9–14 cm depth) and FCOS for each month. The black dashed line denotes the average, and the shaded region represents the interquartile range from 2013 to 2017. The blue circle, red cross and yellow triangle denote the post-thinning years 2020, 2021 and 2022, respectively.Full size imageExamining the environmental variables, June–July 2021 stands out as warmer and drier compared to other years (Fig. 1 and Supplementary Fig. 2). The monthly diel variation of Ta, VPD and soil temperature (Ts) shows that they stayed well above their interquartile range (IQR) from previous years. June 2020 was also warmer and drier than usual, but July and August returned to previous IQR values (Supplementary Fig. 2). PAR remained within average limits during the study period (Fig. 1 and Supplementary Fig. 2) and it peaked during June and July with values ~ 1700 μmol m−2s−1. Ta ranged from −24.7 °C to 30.9 °C during the 3 years of study, and the highest temperatures were measured during June 2021. Similarly, the highest VPD (3.18 kPa) and Ts (16.2 °C) were also measured in the same month. The forest thinning, which happened in January–March of 2020, opened the forest canopy crown, drastically increasing the light availability at the forest floor48. The surface layer soil temperature Ts,humus increased on average by 2 °C in the post-thinning summer months. The precipitation in June and July 2021 was lower than in previous years (Supplementary Fig. 2f), which, in combination with the higher Ts, led to a reduced soil water content (SWC) at the site (Supplementary Fig. 2e). In previous years, SWC peaked in April-May and the minimum occurred in August-September. In 2021, SWC was generally low between June and August, with the minimum in July (Supplementary Fig. 2e). The mean summer SWC was 0.20 ± 0.03 m3m−3 (mean ± standard deviation) during the study period, but the values dropped below 0.14 m3m−3 for a short period in the last 2 weeks of July 2021. However, the water potential at the mineral soil surface dropped to −0.5 MPa on 28 July 2021, which could be considered a mild drought49. In Hyytiälä, the forest generally acts as a sink for COS, with interannual variability dependent on environmental conditions. A reduction in uptake was reported in July 2014, when Ta and VPD were higher5. Similar COS emissions were observed in summer over Harvard Forest under high VPD and Ta conditions37.To understand the ecosystem-level responses and isolate the meteorological influence on FCOS during June and July, FCOS is presented as a function of PAR, Ta and VPD (Fig. 2). The relationships between FCOS and environmental variables during 2020 and 2022 were similar to the pre-thinning years5, showing a decrease of COS uptake with an increase in Ta and VPD, and remained relatively stable under high PAR (>1200 μmol m−2s−1). In 2021, the responses were different, the uptake was lower and net emissions were observed at PAR > 1300 μmol m−2s−1, Ta > 24 °C and VPD > 1.8 kPa. A quadratic function was fitted to the FCOS responses (Supplementary Fig. 3), to determine if the meteorological responses differed significantly between years (p < 0.001). The functions were fitted separately for each year and then combined for all post-thinning years. The fitting coefficients for the quadratic functions are different for 2021 when compared to other years (Supplementary Table 1). The regression metrics (Supplementary Table 1) show that the year-specific model fits were able to explain the variance very well (R2: PAR = 0.94, Ta = 0.93 and VPD = 0.91) compared to one fit for all years (R2: PAR = 0.10, Ta = 0.44 and VPD = 0.39). The improved fit highlights that the interannual variation of the flux responses was significant (p < 0.001).Fig. 2: Correlation between daytime FCOS and driving variables.Binned median correlation between daytime FCOS and environmental variables (a PAR, b Ta and c VPD) during June–July for each year. The half-hourly flux data is sorted equally into ten decile bins, and the shaded regions represent the interquartile range. The number of data points in each bin is indicated in the corresponding panel. The blue circle, red cross, and yellow triangle denote the post-thinning years 2020, 2021 and 2022, respectively.Full size imageReduction in canopy uptake during extreme conditionsThe changes in the meteorological conditions could alter the canopy conductance (GC) and therefore the FCOS. The GC plotted against each season’s low, medium, and high percentiles of PAR from 2020 to 2022 (Fig. 3a) shows that the GC was slightly lower (~12 − 17%) in the 2021 summer. The COS uptake is proportional to the stomatal conductance11, which was observed for all the years, including 2021. The slope of the line fitted against FCOS and GC is similar in all the years (Fig. 3b), with only the intercept being positive in 2021. The intercept shows the FCOS when the canopy conductance is zero, i.e., the contribution from forest floor and soil. The positive intercept points towards an emission source, which offset the non-canopy sinks in 2021.Fig. 3: Variation of canopy conductance with light and daytime FCOS.Sapflow upscaled canopy conductance (GC) and its relation with PAR and FCOS during June–July of each year. a Box-whisker plot for daytime GC in low (10–250 μmol m−2s−1), medium (250–750 μmol m−2s−1) and high (>750 μmol m−2s−1) PAR conditions. The blue, red and yellow-coloured boxes denote the years 2020, 2021 and 2022, respectively. The solid diamonds, circles and squares in each box represent the mean values for 2020, 2021 and 2022, respectively. b Binned median correlation between daytime FCOS and GC (solid lines). The half-hourly data is sorted equally into ten decile bins, and the number of data points in each bin is given. Linear regression between daytime FCOS and GC (dashed lines). The blue circle, red cross and yellow triangle with solid lines denote the years 2020, 2021 and 2022, respectively. The shaded region represents the interquartile range.Full size imageThe influence of meteorological conditions on the reduced COS uptake can also be explored through a meteorologically-driven ecosystem COS flux model. The parameterisation for the daily scale net COS forest sink presented by Vesala et al.5 could be used to quantify the changes in the canopy uptake. The parameterisation considers PAR, Ta, VPD and LAI to estimate COS uptake, reflecting its sensitivity to stomatal conductance (R2 = 0.73, p < 0.001). The parameterisation scheme could closely follow the observed fluxes on a daily scale and effectively capture seasonal variations. In this study, the parameterisation was applied to the Hyytiälä forest stand, accounting for the thinning-related drop in LAI (Supplementary Fig. 4). The average parameterised canopy uptake for summer months was estimated to be −12.4 ± 4.1, −11.2 ± 2.8 and −13.7 ± 3.4 pmol m−2 s−1, for 2020, 2021 and 2022, respectively. The parameterisation generally underestimated the uptake compared to the EC measurements (2020: 34% and 2022: 20%), possibly due to the underrepresentation of forest floor vegetation in the LAI. However, in 2021, the parameterisation overestimated the COS uptake by a factor of two compared to EC measurements. The parameterisation fails to capture the seasonal pattern observed in 2021 (Supplementary Fig. 5). The changes in parameterised fluxes were observed each year, owing to changes in stomatal conductance and, thus, the daily canopy uptake. The average parameterised fluxes for the June-July months demonstrated a small reduction in canopy uptake in 2021 (~1 pmol m−2 s−1) compared to 2020 and 2022.The parameterisation and the changes in the GC both showed that the uptake of COS has undergone slight changes (~5% and ~ 10%, respectively) over the years, which did not explain the observed emissions. When the parameterised COS fluxes are compared with the measured EC daily mean for the 2021 summer (Supplementary Table 2), there was an unaccounted ~ 8 pmol m−2 s−1 COS flux in the budget of this ecosystem.Changes in soil fluxesIn the absence of direct soil chamber measurements in the post-thinning years (2020–22), a depth-resolved diffusion-reaction model (COS Soil Model, COSSM) was used to simulate soil COS fluxes from meteorological and soil physical inputs30. The thinning opened the canopy and warmed the forest floor compared to the pre-thinning years (pre-thinning Ts,humus: 12.3 ± 2.6 °C; post-thinning Ts,humus: 13.5 ± 2.2, 15.5 ± 2.7, 13.7 ± 2.6 °C during 2020, 2021, 2022, respectively) which can significantly influence the soil COS dynamics. The average modelled net soil COS fluxes for June-July were −2.3 ± 0.8, −1.7 ± 1.4 and −2.0 ± 1.0 pmol m−2 s−1, respectively, for 2020, 2021 and 2022. The higher temperature during the 2021 summer led to an increase in the soil COS production rates, thereby reducing the net soil fluxes (Supplementary Fig. 6). The modelled net soil fluxes were still low compared to total ecosystem-level fluxes, which were in the range of −3.7 ± 18.4 pmol m−2 s−1(FCOS, June-July: 2021). It could not account for the total COS emission observed in 2021.The changes in the soil COS fluxes in response to the warmer forest floor alone cannot explain the observed emissions. The COSSM used the soil chamber dataset of Hyytiälä from the growing season of 201518 for fitting the model. The soil chamber measurements during 2015 were on bare soil without a moss or litter layer, and the temperature throughout the measurement was lower than 18.0 °C. Aslan et al.48 showed that in Hyytiälä, light penetration onto the forest floor was much lower in the pre-thinning years (~0.2) compared to post-thinning (~0.4). These factors could contribute to the uncertainty in model-predicted soil COS fluxes. Nevertheless, an increase in soil COS production is most likely in the post-thinning years. At high temperatures and radiation exposure, soils have been found to emit COS14,32,33. Soil dryness could also enhance abiotic COS production14,34. Wheat field soil shifted from consumption to production in the dark at high (>25 °C) temperatures, and sterilised soil shifts from uptake to emissions in the dark measured at a temperature of 19 °C33. Sulfur-abundant agricultural soils emitted COS at ~ 20 °C in dark conditions34. Bryophytes turned from a sink of COS in the dark to a source in light conditions50. Even though other ecosystems have demonstrated COS soil production, it is unlikely that the soil alone could explain the ecosystem-level emissions at Hyytiälä.Spielmann et al.36 have shown that COS emissions from mountain grasslands after harvest hint towards possible COS sources from dead plant matter degradation. Meredith et al.34 found that the microbial cycling of sulfur-containing amino acids, which are COS precursors, was the underlying cause of COS soil fluxes in soil samples from 58 sites. Their study further found that the increase in temperature could explain the apparent light effect, and maximum emissions were observed in boreal peatland soils. The exposure of litter to UV radiation has been found to increase the photodegradation by direct photolysis and also facilitate the microbial degradation by making the organic compounds bioavailable51,52. Solar radiation can also cause thermal and photodegradation of COS precursors analogous to the processes observed in marine ecosystems53,54. Differentiating the direct photolysis and microbial activity is difficult, but UV exposure has consistently enhanced COS emissions or decreased uptake by soil and litter samples, particularly those with high organic matter content55. COS photoproduction from the soil and litter layer was omitted in the COSSM soil model, which could contribute to the observed anomalous emissions.Nighttime fluxesThe light independence of plant CA enzyme and incomplete closure of leaf stomata lead to nighttime canopy uptake of COS11,14. Soil chamber measurements from Hyytiälä have shown that nighttime soil uptake was −2.7 pmol m−2 s−1 during the measurement in the 2015 growing season18. The forest was a nighttime COS sink (−12.4 ± 10.6 pmol m−2 s−1) during the 2013–2017 measurement period. The mean nighttime uptake during 2021 (−3.2 ± 14.4 pmol m−2 s−1) was less than other years (2020: −10.7 ± 11.7 pmol m−2 s−1 and 2022: −9.0 ± 14.4 pmol m−2 s−1), demonstrating reduced ecosystem uptake. Kooijmans et al.11 have shown that the nighttime uptake of COS is controlled by stomatal uptake and reduces with an increase in Ta and VPD. The relation of nighttime FCOS to Ta and VPD (Supplementary Fig. 7) shows this dependence across all years, i.e., reduction of COS uptake as Ta and VPD increase. Even though the median COS nighttime uptake in 2021 was smaller than in other years, uptake still happened at night. The higher average nighttime Ta reduced the stomatal conductance, which would partially explain the reduction in nighttime COS uptake. However, in similar Ta and VPD ranges, the COS uptake was lower in 2021. COS emissions from litter due to the reminiscent heat on the forest floor could not be ruled out, as small emissions were observed at higher temperature conditions at night. Therefore, the emission mechanism could have a thermal component along with a photolytic component that was more active in the daytime.COS emission from a disturbed forest standThe thinning in 2020 disturbed the ecosystem, reduced the GPP by 20% compared to the long-term mean and made it a net source of CO248. The forest stand was able to recover to long-term mean values by 202248. The recovery of the forest was reflected in the response of GPP estimated by partitioning the NEE (GPPNEE), as shown in Fig. 4a. Further, we compare the response of GPP from COS (GPPCOS) in the post-thinning years with the pre-thinning years (Fig. 4b). The responses were noisier than GPPNEE, and the recovery of the forest from the thinning was not visible. The GPPCOS during the year 2022 was lower than in 2020, which is clearly not the case in GPPNEE. Generally, GPPCOS overestimates the GPP compared to GPPNEE3,47, but in 2021, GPPCOS deviated from the normal response of the ecosystem and clearly underestimates GPPNEE.Fig. 4: Comparison of GPP estimated from traditional partitioning and FCOS.a Binned median correlation between daytime GPPNEE and environmental variables (PAR, Ta and VPD) during June-July for each year. The half-hourly data is sorted equally into ten decile bins, and the number of data points in each bin is given. b same as a but for GPPCOS. The black square, blue circle, red cross and yellow triangle denote the years 2013–17, 2020, 2021 and 2022, respectively. The shaded region represents the interquartile range.Full size imageThe disturbance in the boreal forest stand has disrupted the COS budget with an unidentified and potentially significant emission source. In order to estimate the emission in the 2021 summer, the COS uptake by the forest has to be quantified indirectly. The COS uptake by the forest stand can be quantified by inverting the COS-GPP relation (FCOS,GPP)2. Assuming that the forest was the major sink, the difference between the FCOS and COS uptake thus gives the COS emissions (equation (11), Supplementary Fig. 8, Fems:COS,GPP). The mean estimated COS emissions Fems:COS,GPP thus estimated were −6.0 ± 14.4, 23.3 ± 17.5 and 2.7 ± 14.4 pmol m−2 s−1, respectively, in 2020, 2021 and 2022 for the June-July months. This partitioning method for COS had its uncertainties associated with using GPPNEE, LRU and nighttime COS uptake. The emission in 2020 was negative since FCOS,GPP was lower than FCOS, i.e., there was an extra COS sink in 2020 that could not be accounted for by the GPP in that year. Another method to estimate COS uptake is to assume a linear relation between GC and FCOS and use the slope during 2020 and 2022 to estimate the COS uptake in 2021 FCOS,GC. The difference between FCOS,GC and FCOS (equation (12), Supplementary Fig. 8) would give us an estimate of COS emission (Fems:COS,GC: −0.8 ± 12.9, 12.2 ± 14.3 and 0.8 ± 12.5 pmol m−2 s−1 for the years 2020, 2021 and 2022, respectively). FCOS,GC considers canopy-level conductance upscaled from a few tree samples in the forest stand and completely ignores the subcanopy vegetation and soil uptake of COS fluxes (Supplementary Fig. 8). However, FCOS,GC appear closer to FCOS in 2020 and 2022 (Supplementary Table 2). Even though both estimation methods have their own uncertainties, they showed COS emission in 2021 that overwhelmed the average uptake of COS in June–July. The COS atmospheric concentrations (Supplementary Fig. 9) were high during June-July of 2021 (396.6 ± 28.9) compared to the other two years (2020: 281.7 ± 39.4 and 2022: 307.1 ± 37.6). The leaf and soil COS uptake are concentration-dependent39; therefore, the estimated emissions could be underestimated.Despite the aforementioned uncertainties, the main drivers for Fems:COS,GPP and Fems:COS,GC could be clearly identified (Supplementary Fig. 10). The correlation was significant only in the 2021 summer, as the Fems:COS,GPP and Fems:COS,GC are close to zero in 2020 and 2022. The univariate analysis against PAR, Ta, VPD, Ts,humus and surface layer soil water content (SWChumus), shows that Ta and Ts,humus and PAR primarily drive the emissions in both daily and weekly scales. Various combinations of the drivers, avoiding intercorrelation, were used as drivers in the multivariate analysis. A combination of PAR + Ta and PAR + Ts,humus was able to best explain the emissions. It is likely that the emissions were driven by a photothermal component, and we are unsure if it is the temperature that drives the effect of radiation or vice versa.The nature of the emission sources, whether they are occurring in hotspots or spread evenly through the footprint of the EC measurements, could be inferred from a flux-weighted map of positive FCOS (Supplementary Fig. 11) and weighted mean polar plots (Supplementary Fig. 12). Both analyses point towards the unknown COS source distributed heterogeneously within the footprint area. Thinning in the Hyytiälä forest stand has opened up the canopy and doubled the amount of radiation reaching the forest floor (PARff in pre-thinning years: 100.2 ± 112.9 μmol m−2s−1 to 242.8 ± 245.8 μmol m−2s−1 in 2020). The forest floor vegetation recovered steadily in 2021 and 2022, as demonstrated by the forest floor GPP48. The forest floor (Rff) and ecosystem (Reco) respiration fluxes during June were higher in 2021 compared to 2020 and 2022, which was evident from the nighttime NEE fluxes (Supplementary Fig. 13) and modelled respiration (Supplementary Fig. 14). The total respiration for the entire growing season in 2021 was lower than in 2020 or 202248, but the warmer June-July months have seemingly spiked the ecosystem respiration in 2021 (Supplementary Fig. 14). A similar pattern can be found in the forest floor NEE (Supplementary Fig. 14).The presence of S-containing amino acids (SAAs) in soil was found to be a potential precursor of COS34,56. These SAAs could be introduced into the soil through the excess litter left after the thinning. The amount of leaf litter of various species (Supplementary Table 4) over the years and the estimate of S content57,58,59,60 in them can give a rough estimate of available S in the soil. The yearly S-content from litter was 0.41 ± 0.1, 0.46 ± 0.1 and 0.44 ± 0.1 g S m−2 in 2020, 2021 and 2022, respectively, which may drive the observed COS emissions.The COS emission during the senescing period of wheat soil was associated with the mobilisation of SAA and its metabolism14. It should be noted that the surface moss layer in Hyytiälä was considered a COS sink18, due to the consumption of COS by CA enzymes in mosses. The thinning resulted in ~120 g C m−2 decomposable fine-cutting residue, including leaves and needles. Aslan et al.48 estimated that ~50% of the residue already decomposed in 2020–2022 post-thinning years, assuming an exponential decomposition rate61. An initial delay in the decomposition is possible, as the cutting residue was relatively intact and was not incorporated into the litter layer, as it was not compacted by snow right after the thinning in winter 2020. The cutting residue was partially degraded over the following winter and was more susceptible to photothermal degradation over the summer of 2021. The decomposition rate depends on environmental variables such as soil temperature, which was above the seasonal averages, especially in the summer of 2021. The combined effect of the delay in the decomposition of pine needles and the higher temperature possibly resulted in elevated decomposition rates in the 2021 summer. Consequently, the elevated decomposition likely contributed to COS emission from the forest floor.The source of the COS emission observed in summer 2021 can only be speculated at this point. The global COS budget remains unbalanced, and we have demonstrated that the boreal forest could be a source when disturbed. A complete understanding of the COS sources and sinks is necessary for its effective use as a GPP tracer. The continuous observations from the forest floor in a disturbed forest stand could advance our understanding of COS fluxes. Identifying and quantifying this newly observed yet unknown source is required to close the COS budget, reducing a key uncertainty in using COS as a tracer for global photosynthesis. This improvement will help refine carbon cycle models and enhance our ability to monitor and predict ecosystem responses to climate change, with implications for science and policy.MethodsSite description and thinningThe flux measurements from the SMEAR II station in Hyytiälä, Finland (61°51’N, 24°17’E; 181 m a.s.l.) were used in this study. The forest stand is dominated by 50–60-year-old Scots pine (Pinus sylvestris L.) and consists of other tree species, including Norway spruce (Picea abies (L.) Karst.) and deciduous trees (e.g., Betula sp., Populus tremula, and Sorbus aucuparia). Understorey vegetation is characterised by dominant shrub species, notably lingonberry (Vaccinium vitis-idaea L.) and bilberry (Vaccinium myrtillus L.), and prominent mosses such as Schreber’s big red stem moss (Pleurozium schreberi (Brid.) Mitt.)62. In 2018, the basal area was 30.7 m2 ha−1, basal area weighted mean canopy height was 19.9 m, one-sided leaf area index (LAI) 3.9 m2 m−2, and stem density 1141 stems ha−1. The stand was thinned in two phases, i.e., understorey removal and main thinning in March–April 2019 and January–March 2020 (Supplementary Fig. 15), respectively, reducing the basal area by ~ 40%, and foliage by ~ 45%, stem density by ~ 60%48,63. Post-thinning measurements indicated a basal area of 18.0 m2 ha−1, LAI of 2.1 m2 m−2, and stem density of 465 stems ha−1. The species composition remained unchanged after the thinning, as the thinning ratio was almost uniform. Due to canopy opening, light availability almost doubled at the forest floor, and wind speed and daytime air temperature increased at the trunk space in 2020 compared to pre-thinning years48. The daytime duration varied greatly over the months, from ~19.5 h in summer to ~8.3 h in winter. All results were presented in Finnish winter time (UTC + 2). Daytime is defined as the solar elevation angle greater than 0 degrees, and nighttime is less than 0 degrees.EC and other measurementsFCOS were measured during 2020–2022 at the height of 23 m, which consists of an ultrasonic anemometer (uSonic-3 Class A, METEK Meteorologische Messtechnik GmbH, Elmshorn, Germany) for measuring wind speed in three dimensions and sonic temperature; an Aerodyne quantum cascade laser spectrometer (QCLS; Aerodyne Research Inc., Billerica, MA, USA) for measuring COS, CO2, carbon monoxide (CO) and water vapour (H2O) mole fractions. The processing and gap-filling of the COS fluxes were described by Kohonen et al.64. The setup was similar to that reported by Vesala et al.5, which used the data from 2013–17 from Kohonen, K.-M. & Kooijmans, L.65.The EC data were processed using the EddyUH software66 following the procedure recommended by Kohonen et al.64 for COS flux processing. Raw data were despiked so that the difference between subsequent 10 Hz data points was a maximum of 200 ppt for the COS mixing ratio and 5 m s−1 for the vertical wind velocity component (w). The coordinate system was set using a 2D-coordinate rotation, and linear detrending was used to separate the time series into mean and fluctuating components. The time lag of COS was determined by maximising the cross-covariance between CO2 mixing ratio and w. High-frequency spectral losses were corrected using the experimental approach67 according to Mammarella et al.68 so that the response time of CO2 was used for COS spectral correction. Low-frequency correction was done according to Rannik and Vesala69. Finally, only 60% of all the half-hourly fluxes that met the following criteria were accepted: the number of spikes in the 30-minute COS mixing ratio and w less than 100, the second wind rotation angle less than 10 degrees in absolute value, flux stationarity less than 0.3, kurtosis between 1 and 8, and skewness between −2 and 2 (Flag 0 and 1 in Supplementary Fig. 16).The net ecosystem exchange of CO2 was separately measured from above the forest canopy (NEEeco) and at the forest floor (NEEff) using EC. The NEEeco was measured at a height of 27 m above the ground using a closed-path nondispersive infrared sensor (LI-7200, LI-COR Biosciences, USA) paired with an ultrasonic three-dimensional anemometer (HS-50, Gill Ltd, UK) measuring at 10 Hz. This setup is part of the Integrated Carbon Observation System (ICOS) Ecosystem Station network70,71 and the measurements are according to the ICOS recommendations72. (NEEff) was measured at 2.4 m height with an EC system that comprised of a closed-path nondispersive infrared sensor (LI-7200, LI-COR Biosciences, USA) along with a three-dimensional ultrasonic anemometer (model USA-1; METEK GmbH, Elmshorn, Germany). Gas sampling was conducted using an Eaton Synflex tube, with diameters of 6 mm (0.35 m length) and 12.5 mm (1.67 m length), at a flow rate of 12 LPM.The temperature was measured at two heights, 16.8 m and 33.6 m, using Pt100 sensors inside ventilated radiation shields. Similarly, RH was measured at 16.8 and 35 m heights using Rotronic MP102H RH sensors. The average of the measurements at these two heights was used to represent the conditions at the flux measurement height of 23 m. Photosynthetically active radiation (PAR) was measured at 35 m height using a LI-COR LI-190SZ quantum sensor. Ts was measured at five locations in the following depths: organic layer(soil surface), A1 (2–5 cm), B1 horizon (9–14 cm depth), B2 horizon (22–29 cm) and C horizon (42–58 cm) of the forest floor using Philips KTY81-110 temperature sensors. Soil water content (SWC) was also measured using a Delta-T ML3 sensor at 4 locations at above mentioned depths. The all-sided leaf area index (LAI) was measured for a three-day interval using an array of four under-canopy PAR measurement sensors in four different plots by inverting the Beer-Lambert equation (Supplementary Fig. 4). The average of all the sensors was fitted to a bell curve using the scipy.optimize.curve_fit73 package, and the wintertime LAI was interpolated (October-March).Sap flow density was measured with constant heat-dissipation probes74 in eight Scots pine and four silver birch trees. Sensors were installed at breast height on the northern side of the stem, with the reference and heated needles 10 cm apart. The sensors were covered with a reflective aluminium shelter to prevent temperature artefacts. Data were recorded every minute for trees on an RMD680 channel transmitter (Nokeval, Nokeval Oy, Nokia, Finland). Data were averaged over 15-minute periods, and sap flow density was then calculated according to the original calibration74. Maximum temperature difference under zero flow conditions was estimated according to Oishi et al.75 with the following thresholds: nighttime conditions (PAR < 50 μmol m−2 s−1), low vapour pressure deficit (<0.2 kPa) and a stable temperature difference (coefficient of variation < 0.5% over two hours period). When those conditions were not met, the maximum temperature difference under zero-flow conditions was linearly interpolated from neighbouring days.Sapflow upscaled stomatal conductanceThe sap flow measurements from twelve trees (eight Scots pine and four birches) distributed around the forest stand are used to estimate the transpiration and upscale to the ecosystem level. The diameter class technique for upscaling the sap flow observations is utilised, which is based on the assumption that sap flow rates depend only on the diameter at the breast height (DBH) of the tree76,77,78. The DBH distribution for different species of trees per hectare in the sampled area is given in Supplementary Fig. 17. An exponential function is fitted to the relationship between sap flow rates (Qtree) of the sampled trees and their DBH78,79 for Scots pine and birches separately. The coefficients from the exponential function fit are used to obtain the average sap flow for all given tree DBH classes. Further, sap flow for all the DBH classes was aggregated accordingly for sample areas (on a scale of 1 ha) and divided by the average sap flow for sample trees measured within the sample areas. This allowed us to obtain a specific scaling factor (Fs) for sap flow measurements at thinned and control areas, transforming the measurements performed on sample trees to an ecosystem scale. The stand transpiration (TSF) for sample areas was obtained equation (1).$$TSF=frac{{sum }_{i=1}^{n}[{({Q}_{tree})}_{i}cdot {F}_{s}]}{A}$$
    (1)
    The stomatal conductance upscaled from sapflow was calculated by dividing TSF by VPD (equation (2)).$${G}_{C}=frac{TSF}{VPD}$$
    (2)
    Upscaling sap flow measurements from tree level to ecosystem level is challenging and brings uncertainty to achieved results78,80. The comparison of upscaled sap flow measurements, which on a daily or longer time scale represent transpiration with fluxes measured by the EC system, is affected by the footprint dimension and tree distribution within the fetch area. It has also been reported that the uncertainty of upscaled sap flow measurements could also result from the systematic underestimation of sap flux by the sensors (especially thermal dissipation probes, operated according to the constant power principle81,82, but also for the Tissue Heat Balance method78). Upscaling the sap flow measurements is more challenging for mixed forests than monocultures, which also causes uncertainties80. Nevertheless, it is a known and widely used method for assessing transpiration at the ecosystem level based on point sap flow measurement in the tree scale.COS uptake parameterisationThe COS fluxes from the Hyytiälä forest stand were parameterised using the PAR, Ta, VPD, and all-sided LAI from the long time series data shown by Vesala et al.5. The LAI was reduced by almost 40% after the thinning in 2020. The parameterisation of COS flux uptake is given as equation (3).$${F}_{COS}=FPARcdot FScdot FVPDcdot FLAI$$
    (3)
    The four functions for the parameterisation and the parameter values were taken from5, where FPAR is the function of stomatal response to PAR; FS is the phenology of biochemical reactions; FVPD gives the stomatal regulation, and FLAI is the function of foliage and canopy light penetration (Supplementary Fig. 18). The functions and parameterisation description can be found in Vesala et al.5. The parameterised COS fluxes are plotted against the measured EC COS fluxes in Supplementary Fig. 5.Mechanistic soil COS flux modelWe used a depth-resolved diffusion-reaction model (COS Soil Model, COSSM) to simulate soil COS fluxes from meteorological and soil physical inputs30. The model represents both abiotic COS source and enzymatic COS uptake activities and evolves the COS concentration profile prognostically to solve for soil-atmosphere COS fluxes. Here, key improvements have been made for simulating soil COS fluxes at Hyytiälä. Firstly, the model framework has been reimplemented in a differentiable manner using the JAX library in Python83, which facilitates parameter optimisation against soil chamber observations and improves computational efficiency through just-in-time compilation. Secondly, the vertical discretisation of soil layers now follows the Community Land Model version 5 (CLM5)84, with 10 layers of increasing thickness down to 1.5 m depth, fewer than the 25 layers in the original model30. The reduced total number of soil layers and tightly packed soil layers near the surface make simulations faster without aliasing the COS concentration profile.We used chamber measurements of soil COS fluxes at Hyytiälä in 201518 to parameterise the model. Between the two soil chambers in 2015, we used data from soil chamber #1 (SC1) as the training set to optimise COS metabolic parameters and data from soil chamber #2 (SC2) as the test set to assess model generalisability. The parameters that were optimised include COS production capacity (Vmax,P), temperature sensitivity of COS production as indicated by Q10, carbonic anhydrase (CA) COS uptake capacity (Vmax,CA), Gibbs free energy, enthalphy change, and entropy change associated with enzyme activation/deactivation (ΔGa, ΔHd, and ΔSd), and the optimal soil moisture for COS uptake (θopt). The Michaelis constant for CA-mediated COS uptake (Km) is fixed at 0.039 mol m−3 following the experimental value for β-CA7,24. See Supplementary Table 5 for prior and optimised parameter values. The optimised model achieved a root-mean-square error (RMSE) of 1.12 pmol m−2 s−1 on the training set and an RMSE of 1.53 pmol m−2 s−1 on the test set for hourly flux measurements and better performance on daily fluxes (Supplementary Fig. 19). The reduced performance on SC2 data is not surprising given that the root-mean-square deviation between quasi-concurrent SC1 and SC2 flux measurements is 1.50 pmol m−2s−1, reflecting small-scale heterogeneity.We then used meteorological and soil physical observations at Hyytiälä during 2013–2022 to drive the optimised model for half-hourly soil COS flux simulations. June–July soil COS fluxes for 2013–2017 and 2020–2022 are summarised in Supplementary Fig. 6.GPP estimate from COSThe measured EC COS fluxes comprise both vegetation and soil flux components. There were no COS soil chamber measurements during the 2020–2022 period. Hence, average COS soil flux (−2.7 pmol m−2s−1) measured by Sun et al.18 from 2015 in Hyytiälä was used for this analysis. The soil contribution was subtracted from the total EC COS fluxes to estimate the vegetation contribution (FCOS,canopy) and the GPP from COS fluxes (GPPCOS) was estimated using the method described by Sandoval-Soto et al.2 (equation (4)).$$GP{P}_{COS}=frac{-{F}_{COS,canopy}}{LRU}frac{{[C{O}_{2}]}_{a}}{{[COS]}_{a}}$$
    (4)
    where ({[C{O}_{2}]}_{a}) and [COS]a are mixing ratio of CO2 and COS measured by the QCL spectrometer. The leaf relative uptake (LRU) was estimated as a function of PAR (equation (5))3.$$LR{U}_{PAR}=frac{607.26}{PAR}+0.57$$
    (5)
    GPP from partitioning of NEEThe GPPNEE was estimated by traditional NEE flux partitioning (equation (6)) :$$NEE=R-GP{P}_{NLR}$$
    (6)
    where R was estimated from the nighttime NEE fluxes (equation (7)).$$R={R}_{c}{Q}_{10}^{frac{T-10}{10}}$$
    (7)
    where Rc is the respiration at a reference temperature (T = 10 °C) and Q10 is the temperature sensitivity at which the R increases for every 10 °C and T is the driving temperature for the ecosystem. The mean of the soil temperature at 5 cm depth and air temperature at 16.8 m height was found to be a better driving temperature for the entire ecosystem respiration85,86, in the case of the forest floor respiration, only the soil temperature at 5 cm depth is used. The Q10 was 2.0 for the above-canopy fluxes and 2.5 for the forest floor, while Rc was estimated in a moving time window of 11 days (above-canopy fluxes) or 15 days (forest floor fluxes) and seasonal variability ranged from ~ 0.5 to 2.5 μmol m−2 s−1 for above canopy and ~ 0.2 to 1.2 μmol m−2 s−1 for the forest floor. A non-linear regression model for GPP (GPPNLR) is used when the NEE measurements are not available (equation (8)).$$GP{P}_{NLR}=left(sqrt{{(alpha PAR+PA{R}_{max})}^{2}-4Theta PAR{P}_{max}}right)frac{f({T}_{a})}{2Theta }$$
    (8)
    where α and Pmax are fitting parameters that were estimated in a moving time window, f(Ta) is an instantaneous temperature response that reduces the GPP to zero near freezing temperatures (equation (9)). The seasonal variability of Pmax was 0–25 μmol m−2 s−1 and 0–5 μmol m−2 s−1 at the canopy and the forest floor, respectively. α ranged from 0 to 0.04 and from 0 to 0.015 (dimensionless) at the canopy and the forest floor, respectively.$$f({T}_{a})=frac{1}{1+{e}^{2({T}_{0}-{T}_{a})}}$$
    (9)
    where the inflection point is at T0 = − 2 °C87.Estimation of apparent COS emissionsThe GPP-COS relation described by Sandoval-Soto et al.2 can be inverted to estimate the COS uptake by the ecosystem (equation (10)). The GPPNEE was found to follow the natural restoration of the forest stand post-thinning48 (Fig. 4). If there were no emissions in the canopy, the FCOS would have also shown increased uptake as the forest recovered. Therefore, GPPNEE is input in the GPP-COS equation to estimate the COS uptake of the ecosystem (FCOS,GPP) as if the emissions were absent. This method assumes that COS uptake is happening only during photosynthesis and does not consider the COS uptake during the nighttime.$${F}_{COS,GPP}=-GP{P}_{NEE}frac{{[COS]}_{a}}{{[C{O}_{2}]}_{a}}LRU$$
    (10)
    The COS uptake thus estimated was used to find the emission as given in equation (11).$${F}_{ems:COS,GPP}={F}_{COS}-{F}_{COS,GPP}$$
    (11)
    The uptake of COS during the day and night is a direct function of the ecosystem stomatal conductance (GC)11, which can also be used to estimate the COS uptake in the forest stand. Assuming the COS emissions were negligible in the summer of 2020 and 2022, the linear relation between GC and FCOS was estimated (FCOS = −0.21× GC – 12.4) and was used to estimate the canopy uptake of COS (FCOS,GC) in the 2021 summer (Supplementary Table 2). The COS emission was estimated as given in equation (12).$${F}_{ems:COS,GC}={F}_{COS}-{F}_{COS,GC}$$
    (12)

    Data availability

    The code used for analyses can be found in the following GitHub repository https://github.com/t-abin/COS_emission. The COS flux data used in this study can be found at https://doi.org/10.5281/zenodo.17052203. The COSSM model version for simulating soil COS fluxes, along with the required input data, is available at https://doi.org/10.5281/zenodo.17555492. The flux data from the 27m tower (NEE) used in this study can be found at https://doi.org/10.23729/b4575245-5612-4874-a33b-541b215b1a10, and the environmental data at https://doi.org/10.23729/23dd00b2-b9d7-467a-9cee-b4a122486039. Most recent versions of the fluxes and the environmental data can be downloaded from https://smear.avaa.csc.fi.
    ReferencesBerry, J. et al. A coupled model of the global cycles of carbonyl sulfide and CO2 a possible new window on the carbon cycle. J. Geophys. Res. Biogeosci. 118, 842–852 (2013).Article 

    Google Scholar 
    Sandoval-Soto, L. et al. Global uptake of carbonyl sulfide (COS) by terrestrial vegetation: estimates corrected by deposition velocities normalized to the uptake of carbon dioxide (CO2). Biogeosciences 2, 125–132 (2005).Article 

    Google Scholar 
    Kooijmans, L. M. J. et al. Influences of light and humidity on carbonyl sulfide-based estimates of photosynthesis. Proc. Natl. Acad. Sci. USA 116, 2470–2475 (2019).Article 

    Google Scholar 
    Whelan, M. E. et al. Reviews and syntheses: carbonyl sulfide as a multi-scale tracer for carbon and water cycles. Biogeosciences 15, 3625–3657 (2018).Article 

    Google Scholar 
    Vesala, T. et al. Long-term fluxes of carbonyl sulfide and their seasonality and interannual variability in a boreal forest. Atmos. Chem. Phys. 22, 2569–2584 (2022).Article 

    Google Scholar 
    Protoschill-Krebs, G. & Kesselmeier, J. Enzymatic pathways for the consumption of carbonyl sulphide (COS) by higher plants*. Botanica Acta 105, 206–212 (1992).Article 

    Google Scholar 
    Protoschill-Krebs, G., Wilhelm, C. & Kesselmeier, J. Consumption of carbonyl sulphide (COS) by higher plant carbonic anhydrase (CA). Atmos. Environ. 30, 3151–3156 (1996).Article 

    Google Scholar 
    DiMario, R. J., Clayton, H., Mukherjee, A., Ludwig, M. & Moroney, J. V. Plant carbonic anhydrases: structures, locations, evolution, and physiological roles. Mol. Plant 10, 30–46 (2017).Article 

    Google Scholar 
    Schenk, S., Kesselmeier, J. & Anders, E. How does the exchange of one oxygen atom with sulfur affect the catalytic cycle of carbonic anhydrase?. Chem. A Eur. J. 10, 3091–3105 (2004).Article 

    Google Scholar 
    Stimler, K., Montzka, S. A., Berry, J. A., Rudich, Y. & Yakir, D. Relationships between carbonyl sulfide (cos) and co2 during leaf gas exchange. N. Phytol. 186, 869–878 (2010).Article 

    Google Scholar 
    Kooijmans, L. M. J. et al. Canopy uptake dominates nighttime carbonyl sulfide fluxes in a boreal forest. Atmos. Chem. Phys. 17, 11453–11465 (2017).Article 

    Google Scholar 
    Caird, M. A., Richards, J. H. & Donovan, L. A. Nighttime stomatal conductance and transpiration in C3 and C4 plants. Plant Physiol. 143, 4–10 (2007).Article 

    Google Scholar 
    Wehr, R. et al. Dynamics of canopy stomatal conductance, transpiration, and evaporation in a temperate deciduous forest, validated by carbonyl sulfide uptake. Biogeosciences 14, 389–401 (2017).Article 

    Google Scholar 
    Maseyk, K. et al. Sources and sinks of carbonyl sulfide in an agricultural field in the southern great plains. Proc. Natl. Acad. Sci. USA 111, 9064–9069 (2014).Article 

    Google Scholar 
    White, M. L. et al. Carbonyl sulfide exchange in a temperate loblolly pine forest grown under ambient and elevated CO2. Atmos. Chem. Phys. 10, 547–561 (2010).Article 

    Google Scholar 
    Seibt, U., Kesselmeier, J., Sandoval-Soto, L., Kuhn, U. & Berry, J. A. A kinetic analysis of leaf uptake of cos and its relation to transpiration, photosynthesis and carbon isotope fractionation. Biogeosciences 7, 333–341 (2010).Article 

    Google Scholar 
    Yang, F., Qubaja, R., Tatarinov, F., Rotenberg, E. & Yakir, D. Assessing canopy performance using carbonyl sulfide measurements. Glob. Change Biol. 24, 3486–3498 (2018).Article 

    Google Scholar 
    Sun, W. et al. Soil fluxes of carbonyl sulfide (COS), carbon monoxide, and carbon dioxide in a boreal forest in southern finland. Atmos. Chem. Phys. 18, 1363–1378 (2018).Article 

    Google Scholar 
    Sauze, J. et al. The interaction of soil phototrophs and fungi with pH and their impact on soil CO2,CO18O and OCS exchange. Soil Biol. Biochem. 115, 371–382 (2017).Article 

    Google Scholar 
    Meredith, L. K. et al. Soil exchange rates of COS and CO18o differ with the diversity of microbial communities and their carbonic anhydrase enzymes. ISME J. 13, 290–300 (2019).Article 

    Google Scholar 
    Whelan, M. E. et al. Carbonyl sulfide exchange in soils for better estimates of ecosystem carbon uptake. Atmos. Chem. Phys. 16, 3711–3726 (2016).Article 

    Google Scholar 
    Van Diest, H. & Kesselmeier, J. Soil atmosphere exchange of carbonyl sulfide (COS) regulated by diffusivity depending on water-filled pore space. Biogeosciences 5, 475–483 (2008).Article 

    Google Scholar 
    Kesselmeier, J., Teusch, N. & Kuhn, U. Controlling variables for the uptake of atmospheric carbonyl sulfide by soil. J. Geophys. Res. Atmos. 104, 11577–11584 (1999).Article 

    Google Scholar 
    Ogée, J. et al. A new mechanistic framework to predict OCS fluxes from soils. Biogeosciences 13, 2221–2240 (2016).Article 

    Google Scholar 
    Launois, T., Peylin, P., Belviso, S. & Poulter, B. A new model of the global biogeochemical cycle of carbonyl sulfide – part 2: use of carbonyl sulfide to constrain gross primary productivity in current vegetation models. Atmos. Chem. Phys. 15, 9285–9312 (2015).Article 

    Google Scholar 
    Whelan, M. E., Min, D.-H. & Rhew, R. C. Salt marsh vegetation as a carbonyl sulfide (COS) source to the atmosphere. Atmos. Environ. 73, 131–137 (2013).Article 

    Google Scholar 
    Li, X.,Liu,J. & Yang, J. Variation of H2S and COS emission fluxes from calamagrostis angustifolia wetlands in sanjiang plain, northeast china. Atmos. Environ. 40, 6303–6312 (2006).Article 

    Google Scholar 
    Caron, F. & Kramer, J. R. Formation of volatile sulfides in freshwater environments. Sci. Total Environ. 153, 177–194 (1994).Article 

    Google Scholar 
    Kaisermann, A., Jones, S. P., Wohl, S., Ogée, J. & Wingate, L. Nitrogen fertilization reduces the capacity of soils to take up atmospheric carbonyl sulphide. Soil Syt. 2, 62 (2018).Article 

    Google Scholar 
    Sun, W., Maseyk, K., Lett, C. & Seibt, U. A soil diffusion-reaction model for surface COS flux: COSSM v1. Geosci. Model Dev. 8, 3055–3070 (2015).Article 

    Google Scholar 
    Yi, Z. et al. Soil uptake of carbonyl sulfide in subtropical forests with different successional stages in south china. J. Geophys. Res. Atmos. 112, D08302 (2007).Article 

    Google Scholar 
    Kitz, F. et al. In situ soil COS exchange of a temperate mountain grassland under simulated drought. Oecologia 183, 851–860 (2017).Article 

    Google Scholar 
    Whelan, M. E. & Rhew, R. C. Carbonyl sulfide produced by abiotic thermal and photodegradation of soil organic matter from wheat field substrate. J. Geophys. Res. Biogeosci. 120, 54–62 (2015).Article 

    Google Scholar 
    Meredith, L. K. et al. Coupled biological and abiotic mechanisms driving carbonyl sulfide production in soils. Soil Syt. 2, 37 (2018).Article 

    Google Scholar 
    Kitz, F. et al. Soil carbonyl sulfide exchange in relation to microbial community composition: insights from a managed grassland soil amendment experiment. Soil Biol. Biochem. 135, 28–37 (2019).Article 

    Google Scholar 
    Spielmann, F. M., Hammerle, A., Kitz, F., Gerdel, K. & Wohlfahrt, G. Seasonal dynamics of the COS and CO2 exchange of a managed temperate grassland. Biogeosciences 17, 4281–4295 (2020).Article 

    Google Scholar 
    Commane, R. et al. Seasonal fluxes of carbonyl sulfide in a midlatitude forest. Proc. Natl. Acad. Sci. USA 112, 14162–14167 (2015).Article 

    Google Scholar 
    Maignan, F. et al. Carbonyl sulfide: comparing a mechanistic representation of the vegetation uptake in a land surface model and the leaf relative uptake approach. Biogeosciences 18, 2917–2955 (2021).Article 

    Google Scholar 
    Kooijmans, L. M. J. et al. Evaluation of carbonyl sulfide biosphere exchange in the simple biosphere model (sib4). Biogeosciences 18, 6547–6565 (2021).Article 

    Google Scholar 
    Abadie, C. et al. Carbon and water fluxes of the boreal evergreen needleleaf forest biome constrained by assimilating ecosystem carbonyl sulfide flux observations. J. Geophys. Res. Biogeosci. 128, e2023JG007407 (2023).Article 

    Google Scholar 
    Lennartz, S. T. et al. Direct oceanic emissions unlikely to account for the missing source of atmospheric carbonyl sulfide. Atmos. Chem. Phys. 17, 385–402 (2017).Article 

    Google Scholar 
    Ma, J. et al. Inverse modelling of carbonyl sulfide: implementation, evaluation and implications for the global budget. Atmos. Chem. Phys. 21, 3507–3529 (2021).Article 

    Google Scholar 
    Remaud, M. et al. Plant gross primary production, plant respiration and carbonyl sulfide emissions over the globe inferred by atmospheric inverse modelling. Atmos. Chem. Phys. 22, 2525–2552 (2022).Article 

    Google Scholar 
    Cartwright, M. P. et al. Constraining the budget of atmospheric carbonyl sulfide using a 3-d chemical transport model. Atmos. Chem. Phys. 23, 10035–10056 (2023).Article 

    Google Scholar 
    de Vries, A. et al. The contribution of boreal wetlands to the northern hemisphere carbonyl sulfide sink. Geophys. Res. Lett. 52, e2024GL112858 (2025).Article 

    Google Scholar 
    Asaf, D. et al. Ecosystem photosynthesis inferred from measurements of carbonyl sulphide flux. Nat. Geosci. 6, 186–190 (2013).Article 

    Google Scholar 
    Kohonen, K.-M. et al. Intercomparison of methods to estimate gross primary production based on CO2 and COS flux measurements. Biogeosciences 19, 4067–4088 (2022).Article 

    Google Scholar 
    Aslan, T. et al. Thinning turned boreal forest to a temporary carbon source – short term effects of partial harvest on carbon dioxide and water vapor fluxes. Agric. For. Meteorol. 353, 110061 (2024).Article 

    Google Scholar 
    Duursma, R. A. et al. Predicting the decline in daily maximum transpiration rate of two pine stands during drought based on constant minimum leaf water potential and plant hydraulic conductance. Tree Physiol. 28, 265–276 (2008).Article 

    Google Scholar 
    Gimeno, T. E. et al. Bryophyte gas-exchange dynamics along varying hydration status reveal a significant carbonyl sulphide (cos) sink in the dark and cos source in the light. N. Phytol. 215, 965–976 (2017).Article 

    Google Scholar 
    Baker, N. R. & Allison, S. D. Ultraviolet photodegradation facilitates microbial litter decomposition in a mediterranean climate. Ecology 96, 1994–2003 (2015).Article 

    Google Scholar 
    King, J. Y., Brandt, L. A. & Adair, E. C. Shedding light on plant litter decomposition: advances, implications and new directions in understanding the role of photodegradation. Biogeochemistry 111, 57–81 (2012).Article 

    Google Scholar 
    Zepp, R. G. & Andreae, M. O. Factors affecting the photochemical production of carbonyl sulfide in seawater. Geophys. Res. Lett. 21, 2813–2816 (1994).Article 

    Google Scholar 
    Uher, G. & Andreae, M. Photochemical production of carbonyl sulfide in north sea water: a process study. Limnol. Oceanogr. 42, 432–442 (1997).Article 

    Google Scholar 
    Kitz, F. et al. Soil cos exchange: a comparison of three european ecosystems. Glob. Biogeochem. Cycles 34, e2019GB006202 (2020).Article 

    Google Scholar 
    Rennenberg, H.The Significance of Higher Plants in the Emission of Sulfur Compounds from Terrestrial Ecosystems. In Trace Gas Emissions by Plants (eds. Sharkey, T. D., Holland, E. A. & Mooney, H. A.) Physiological Ecology, 217–260 (Academic Press, San Diego, 1991).Mammarella, I. et al. Etc l2 archive from hyytiala, 2018–2024 https://hdl.handle.net/11676/oD2Pme94V4Cl6dmxyZDKuCeq (2025).Likus-Cieślik, J. & Pietrzykowski, M. Vegetation development and nutrients supply of trees in habitats with high sulfur concentration in reclaimed former sulfur mines jeziórko (southern poland). Environ. Sci. Pollut. Res. 24, 20556–20566 (2017).Article 

    Google Scholar 
    Giertych, M. J., de Temmerman, L. O. & Rachwal, L. Distribution of elements along the length of scots pine needles in a heavily polluted and a control environment. Tree Physiol. 17, 697–703 (1997).Article 

    Google Scholar 
    Zethof, J. H. T., Julich, S., Feger, K.-H. & Julich, D. Legacy effect of 25 years reduced atmospheric sulphur deposition on spruce tree nutrition. J. Plant Nutr. Soil Sci. 187, 834–843 (2024).Article 

    Google Scholar 
    Olson, J. S. Energy storage and the balance of producers and decomposers in ecological systems. Ecology 44, 322–331 (1963).Article 

    Google Scholar 
    Kolari, P. et al. Hyytiälä SMEAR II site characteristics https://doi.org/10.5281/zenodo.5909681 (2022).Aalto, J. et al. Hyytiälä SMEAR II forest year 2020 thinning tree and carbon inventory data https://doi.org/10.5281/zenodo.8138946 (2023).Kohonen, K.-M. et al. Towards standardized processing of eddy covariance flux measurements of carbonyl sulfide. Atmos. Meas. Tech. 13, 3957–3975 (2020).Article 

    Google Scholar 
    Kohonen, K.-M. & Kooijmans, L. Dataset for “Long-term fluxes of carbonyl sulfide and their seasonality and interannual variability in a boreal forest” https://doi.org/10.5281/zenodo.5906705 (2022).Mammarella, I., Peltola, O., Nordbo, A., Järvi, L. & Rannik, U. Quantifying the uncertainty of eddy covariance fluxes due to the use of different software packages and combinations of processing steps in two contrasting ecosystems. Atmos. Meas. Tech. 9, 4915–4933 (2016).Article 

    Google Scholar 
    Aubinet, M. et al. Estimates of the annual net carbon and water exchange of forests: The EUROFLUX methodology. In Advances in Ecological Research (Fitter, A. & Raffaelli, D. eds) Vol. 30, 113–175 (Academic Press, 1999).Mammarella, I. et al. Relative humidity effect on the high-frequency attenuation of water vapor flux measured by a closed-path eddy covariance system. J. Atmos. Ocean. Technol. 26, 1856 – 1866 (2009).Article 

    Google Scholar 
    Rannik, Ü. & Vesala, T. Autoregressive filtering versus linear detrending in estimation of fluxes by the eddy covariance method. Bound.-Layer. Meteorol. 91, 259–280 (1999).Article 

    Google Scholar 
    Franz, D. et al. Towards long-term standardised carbon and greenhouse gas observations for monitoring europe’s terrestrial ecosystems: a review. Int. Agrophys. 32, 439–455 (2018).Article 

    Google Scholar 
    Heiskanen, J. et al. The integrated carbon observation system in europe. Bull. Am. Meteorol. Soc. 103, E855 – E872 (2022).Article 

    Google Scholar 
    Rebmann, C. et al. Icos eddy covariance flux-station site setup: a review. Int. Agrophys. 32, 471–494 (2018).Article 

    Google Scholar 
    Virtanen, P. et al. Scipy 1.0: fundamental algorithms for scientific computing in python. Nat. Methods 17, 261–272 (2020).Article 

    Google Scholar 
    Granier, A. Une nouvelle méthode pour la mesure du flux de sève brute dans le tronc des arbres. In Annales des Sciences Forestières Vol. 42, 193–200 (EDP Sciences, 1985).Oishi, A. C., Oren, R. & Stoy, P. C. Estimating components of forest evapotranspiration: a footprint approach for scaling sap flux measurements. Agric. For. Meteorol. 148, 1719–1732 (2008).Article 

    Google Scholar 
    Cermák, J. & Kučera, J. Transpiration of mature stands of spruce (picea abies (l.) karst.) as estimated by the tree-trunk heat balance method. In Forest Hydrology and Watershed Management (eds. Swanson, R, W. P. & Bernier P) 311–317 (Wallingford, UK, 1987).Cermák, J. & Kučera, J. Scaling up transpiration data between trees, stands and watersheds. Silva Carelica 15, 101–120 (1990).
    Google Scholar 
    Dukat, P. et al. Scots pine responses to drought investigated with eddy covariance and sap flow methods. Eur. J. For. Res. 142, 671–690 (2023).Article 

    Google Scholar 
    Cristiano, P. M. et al. Evapotranspiration of subtropical forests and tree plantations: a comparative analysis at different temporal and spatial scales. Agric. For. Meteorol. 203, 96–106 (2015).Article 

    Google Scholar 
    Rabbel, I., Bogena, H., Neuwirth, B. & Diekkrüger, B. Using sap flow data to parameterize the Feddes water stress model for Norway spruce. Water 10, 279 (2018).Wilson, K. B., Hanson, P. J., Mulholland, P. J., Baldocchi, D. D. & Wullschleger, S. D. A comparison of methods for determining forest evapotranspiration and its components: sap-flow, soil water budget, eddy covariance and catchment water balance. Agric. For. Meteorol. 106, 153–168 (2001).Article 

    Google Scholar 
    Granier, A. Evaluation of transpiration in a Douglas-fir stand by means of sap flow measurements. Tree Physiol. 3, 309–320 (1987).Article 

    Google Scholar 
    Frostig, R., Johnson, M. & Leary, C. Compiling Machine Learning Programs via High-Level Tracing. In SysML 18 https://mlsys.org/Conferences/doc/2018/146.pdf (Stanford, CA, USA, 2018).Lawrence, D. M. et al. The community land model version 5: Description of new features, benchmarking, and impact of forcing uncertainty. J. Adv. Model. Earth Syst. 11, 4245–4287 (2019).Article 

    Google Scholar 
    Kolari, P. et al. Co2 exchange and component co2 fluxes of a boreal scots pine forest. Boreal Environ. Res. 14, 761–783 (2009).
    Google Scholar 
    Lasslop, G. et al. On the choice of the driving temperature for eddy-covariance carbon dioxide flux partitioning. Biogeosciences 9, 5243–5259 (2012).Article 

    Google Scholar 
    Kolari, P. et al. Field and controlled environment measurements show strong seasonal acclimation in photosynthesis and respiration potential in boreal scots pine. Front. Plant Sci. 5, 717 (2014).Article 

    Google Scholar 
    Download referencesAcknowledgementsThe authors are grateful to Sirpa Rantanen, Elmeri Putkinen, and other technical support staff at the Hyytiälä forest station for their expertise and hard work in ensuring the smooth operation and maintenance of the measurement equipment. This study has received financial support from the University of Helsinki via Integrated Carbon Observation System – Hyytiälä (ICOS-HY), the Research Council of Finland project N-PERM (Grant No. 341348), and project ForClimate (Grant No. 347780), and the EU Horizon Europe-Framework Programme for Research and Innovation (Grant No. 101056921-GreenFeedBack).Author informationAuthors and AffiliationsInstitute for Atmospheric and Earth System Research (INAR)/Physics, Faculty of Science, University of Helsinki, Helsinki, FinlandAbin Thomas, Asta Laasonen, Paulina Dukat, Yann Salmon, Roderick Dewar, Pasi Kolari, Timo Vesala & Ivan MammarellaDepartment of Environmental Systems Science, Institute of Agricultural Sciences, ETH Zurich, Zurich, SwitzerlandKukka-Maaria KohonenFinnish Meteorological Institute, Helsinki, FinlandToprak AslanInstitute of Landscape Ecology, University of Münster, Münster, GermanyToprak AslanDepartment of Global Ecology, Carnegie Institution for Science, Stanford, California, USAWu SunInstitute for Atmospheric and Earth System Research (INAR)/Forest Sciences, Faculty of Agriculture and Forestry, University of Helsinki, Helsinki, FinlandPaulina Dukat, Yann Salmon & Timo VesalaPlant Sciences Division, Research School of Biology, The Australian National University, Canberra, ACT, AustraliaRoderick DewarSchool of Environment, Earth and Ecosystem Sciences, The Open University, Milton Keynes, UKKadmiel MaseykAuthorsAbin ThomasView author publicationsSearch author on:PubMed Google ScholarAsta LaasonenView author publicationsSearch author on:PubMed Google ScholarKukka-Maaria KohonenView author publicationsSearch author on:PubMed Google ScholarToprak AslanView author publicationsSearch author on:PubMed Google ScholarWu SunView author publicationsSearch author on:PubMed Google ScholarPaulina DukatView author publicationsSearch author on:PubMed Google ScholarYann SalmonView author publicationsSearch author on:PubMed Google ScholarRoderick DewarView author publicationsSearch author on:PubMed Google ScholarPasi KolariView author publicationsSearch author on:PubMed Google ScholarKadmiel MaseykView author publicationsSearch author on:PubMed Google ScholarTimo VesalaView author publicationsSearch author on:PubMed Google ScholarIvan MammarellaView author publicationsSearch author on:PubMed Google ScholarContributionsA.T., I.M., and T.V. conceived and designed the study. AT wrote the main manuscript and created figures 1-4. A.L., T.A. and K.M.K. collected, processed and finalised the COS eddy covariance data. Y.S. provided the sapflow measurements data. P.K. managed and facilitated the data flow and processed the NEE eddy covariance and environmental datasets. P.D. helped with upscaling the sapflow measurements and writing relevant parts in the manuscript. W.S. ran the differentiable COSSM model, performed analysis and contributed to the section on changes in soil fluxes. K.M.K., K.M., R.D., T.V. and I.M. helped with the conceptual framework of the manuscript and editing of the final version. All authors reviewed the manuscript.Corresponding authorCorrespondence to
    Abin Thomas.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 informationSupplementary InformationRights 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 articleThomas, A., Laasonen, A., Kohonen, KM. et al. Examining anomalous summer carbonyl sulfide emissions in a boreal forest after thinning.
    npj Clim Atmos Sci 9, 1 (2026). https://doi.org/10.1038/s41612-025-01272-wDownload citationReceived: 05 September 2025Accepted: 11 November 2025Published: 02 January 2026Version of record: 02 January 2026DOI: https://doi.org/10.1038/s41612-025-01272-wShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    A green patina unlocking cities’ potential to adapt to climate change

    Not only are increasing numbers of people in cities suffering from climate change, hotter temperatures, and drought, but so is green infrastructure, which is inherently there for recreation and cooling. This Comment argues that the entire green patina of the city is needed to adapt effectively.

    Access through your institution

    Buy or subscribe

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

    Access options

    Access through your institution

    Subscribe to this journal

    Receive 12 digital issues and online access to articles

    $119.00 per year
    only $9.92 per issue

    Learn more

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

    Additional access options:

    Log in

    Learn about institutional subscriptions

    Read our FAQs

    Contact customer support

    ReferencesLin, B. B. et al. Lancet Planet. Health 5, e479–e486 (2021).Article 

    Google Scholar 
    Egerer, M. et al. npj Urban Sustain 1, 22 (2021).Article 

    Google Scholar 
    Wang, S. et al. Nat. Clim. Chang. 15, 497–504 (2025).Article 

    Google Scholar 
    Pan, H. et al. Nat. Clim. Chang. 13, 862–870 (2023).Article 

    Google Scholar 
    Ignatieva, M., Haase, D., Dushkova, D. & Haase, A. Land 9, 73 (2020).Article 

    Google Scholar 
    Chancel, L. et al. Nat. Clim. Chang. 15, 364–374 (2025).Article 

    Google Scholar 
    Haase, D. & Hellwig, R. Trees For. People 8, 100252 (2022).Article 

    Google Scholar 
    Leisenheimer, L., Wellmann, T., Jänicke, C. & Haase, D. Ecol. Inform. 82, (2024).Wellmann, T. et al. Ambio 52, 489–507 (2023).Article 

    Google Scholar 
    McDonald, R. I. et al. Nat. Sustain. 3, 16–24 (2020).Article 

    Google Scholar 
    Cano, Á. et al. Nat. Ecol. Evol. 9, 214–224 (2025).Article 

    Google Scholar 
    Forseth, I. N. & Innis, A. F. Crit. Rev. Plant Sci. 23, 401–413 (2004).Article 

    Google Scholar 
    Pinkert, S., Farwig, N., Kawahara, A. Y. & Jetz, W. Nat. Ecol. Evol. 9, 789–800 (2025).Article 

    Google Scholar 
    Elmqvist, T. et al. npj Urban Sustain 1, 6 (2021).Article 

    Google Scholar 
    McPhearson, T. et al. npj Urban Sustain 1, 5 (2021).Article 

    Google Scholar 
    Download referencesAcknowledgementsThis research was funded by the NaturaConnect (Designing a Resilient and Coherent Trans-European Network for Nature and People) Horizon 2020 project (contract no. 101060429).Author informationAuthors and AffiliationsDepartment of Geography, Humboldt University Berlin, Berlin, GermanyDagmar HaaseDepartment of Computational Landscape Ecology, Helmholtz Centre for Environmental Research – UFZ, Leipzig, GermanyDagmar HaaseAuthorsDagmar HaaseView author publicationsSearch author on:PubMed Google ScholarCorresponding authorCorrespondence to
    Dagmar Haase.Ethics declarations

    Competing interests
    The author declares no competing interests.

    Peer review

    Peer review information
    Nature Cities thanks Xuemei Bai, Durwood Zaelke and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

    Rights and permissionsReprints and permissionsAbout this articleCite this articleHaase, D. A green patina unlocking cities’ potential to adapt to climate change.
    Nat Cities (2026). https://doi.org/10.1038/s44284-025-00358-zDownload citationPublished: 02 January 2026Version of record: 02 January 2026DOI: https://doi.org/10.1038/s44284-025-00358-zShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More