More stories

  • in

    Collective behaviour can stabilize ecosystems

    1.Chesson, P. General theory of competitive coexistence in spatially-varying environments. Theor. Popul. Biol. 58, 211–237 (2000).2.Hubbell, S. P. The Unified Neutral Theory of Biodiversity and Biogeography (Princeton Univ. Press, 2001).3.Ellner, S. P., Snyder, R. E., Adler, P. B. & Hooker, G. An expanded modern coexistence theory for empirical applications. Ecol. Lett. 22, 3–18 (2019).4.Rosenzweig, M. L. Paradox of enrichment: destabilization of exploitation ecosystems in ecological time. Science 171, 385–387 (1971).5.Costantino, R. F., Cushing, J. M., Dennis, B. & Desharnais, R. A. Experimentally induced transitions in the dynamic behaviour of insect populations. Nature 375, 227–230 (1995).6.Fussmann, G. F., Ellner, S. P., Shertzer, K. W. & Hairston, N. G. Jr. Crossing the Hopf bifurcation in a live predator-prey system. Science 290, 1358–1360 (2000).7.Dalziel, B. D. et al. Persistent chaos of measles epidemics in the prevaccination United States caused by a small change in seasonal transmission patterns. PLoS Comput. Biol. 12, e1004655 (2016).8.Darwin, C. On the Origin of Species by Means of Natural Selection, or The Preservation of Favoured Races in the Struggle for Life (John Murray, 1859).9.Gause, G. F. Experimental analysis of Vito Volterra’s mathematical theory of the struggle for existence. Science 79, 16–17 (1934).10.Hutchinson, G. E. The paradox of the plankton. Am. Nat. 95, 137–145 (1961).Article 

    Google Scholar 
    11.Chesson, P. Multispecies competition in variable environments. Theor. Popul. Biol. 45, 227–276 (1994).Article 

    Google Scholar 
    12.McCann, K., Hastings, A. & Huxel, G. R. Weak trophic interactions and the balance of nature. Nature 395, 794–798 (1998).13.Rooney, N., McCann, K., Gellner, G. & Moore, J. C. Structural asymmetry and the stability of diverse food webs. Nature 442, 265–269 (2006).14.Coyte, K. Z., Schluter, J. & Foster, K. R. The ecology of the microbiome: networks, competition, and stability. Science 350, 663–666 (2015).15.May, R. M. Host-parasitoid systems in patchy environments: a phenomenological model. J. Anim. Ecol. 47, 833–844 (1978).Article 

    Google Scholar 
    16.Briggs, C. J. & Hoopes, M. F. Stabilizing effects in spatial parasitoid–host and predator–prey models: a review. Theor. Popul. Biol. 65, 299–315 (2004).17.Vicsek, T. & Zafeiris, A. Collective motion. Phys. Rep. 517, 71–140 (2012).Article 

    Google Scholar 
    18.Berdahl, A., Torney, C. J., Ioannou, C. C., Faria, J. J. & Couzin, I. D. Emergent sensing of complex environments by mobile animal groups. Science 339, 574–576 (2013).19.Nagy, M., Akos, Z., Biro, D. & Vicsek, T. Hierarchical group dynamics in pigeon flocks. Nature 464, 890–893 (2010).20.Dalziel, B. D., Corre, M. L., Côté, S. D. & Ellner, S. P. Detecting collective behaviour in animal relocation data, with application to migrating caribou. Methods Ecol. Evol. 7, 30–41 (2015).Article 

    Google Scholar 
    21.Torney, C. J. et al. Inferring the rules of social interaction in migrating caribou. Phil. Trans. R. Soc. B 373, 20170385 (2018).22.Fryxell, J. M., Mosser, A., Sinclair, A. R. E. & Packer, C. Group formation stabilizes predator–prey dynamics. Nature 449, 1041–1043 (2007).23.Vicsek, T., Czirók, A., Ben-Jacob, E., Cohen, I. & Shochet, O. Novel type of phase transition in a system of self-driven particles. Phys. Rev. Lett. 75, 1226 (1995).CAS 
    Article 

    Google Scholar 
    24.Buhl, J. et al. From disorder to order in marching locusts. Science 312, 1402–1406 (2006).25.King, A. J., Fehlmann, G., Biro, D., Ward, A. J. & Fürtbauer, I. Re-wilding collective behaviour: an ecological perspective. Trends Ecol. Evol. 33, 347–357 (2018).26.Sumpter, D. J. T. Collective Animal Behavior (Princeton Univ. Press, 2010).27.Guttal, V. & Couzin, I. D. Social interactions, information use, and the evolution of collective migration. Proc. Natl Acad. Sci. USA 107, 16172–16177 (2010).CAS 
    Article 

    Google Scholar 
    28.Barbier, M. & Watson, J. R. The spatial dynamics of predators and the benefits and costs of sharing information. PLoS Comput. Biol. 12, e1005147 (2016).29.Lotka, A. J. Analytical note on certain rhythmic relations in organic systems. Proc. Natl Acad. Sci. USA 6, 410–415 (1920).CAS 
    Article 

    Google Scholar 
    30.Rosenzweig, M. L. & MacArthur, R. H. Graphical representation and stability conditions of predator-prey interactions. Am. Nat. 97, 209–223 (1963).Article 

    Google Scholar 
    31.Couzin, I. D., Krause, J., James, R., Ruxton, G. D. & Franks, N. R. Collective memory and spatial sorting in animal groups. J. Theor. Biol. 218, 1–11 (2002).32.Couzin, I. D., Krause, J., Franks, N. R. & Levin, S. A. Effective leadership and decision-making in animal groups on the move. Nature 433, 513–516 (2005).33.MacArthur, R. H. Population ecology of some warblers of northeastern coniferous forests. Ecology 39, 599–619 (1958).Article 

    Google Scholar 
    34.Dalziel, B. D., Thomann, E., Medlock, J. & De Leenheer, P. Global analysis of a predator-prey model with variable predator search rate. J. Math. Biol. 81, 159–183 (2020).35.Lukas, D. & Clutton-Brock, T. Social complexity and kinship in animal societies. Ecol. Lett. 21, 1129–1134 (2018).36.Purves, D. W., Lichstein, J. W., Strigul, N. & Pacala, S. W. Predicting and understanding forest dynamics using a simple tractable model. Proc. Natl Acad. Sci. USA 105, 17018–17022 (2008).CAS 
    Article 

    Google Scholar 
    37.Dalziel, B. D. et al. Urbanization and humidity shape the intensity of influenza epidemics in U.S. cities. Science 362, 75–79 (2018).38.Monk, C. T. et al. How ecology shapes exploitation: a framework to predict the behavioural response of human and animal foragers along exploration-exploitation trade-offs. Ecol. Lett. 21, 779–793 (2018).39.Hutchins, D. A. & Fu, F. Microorganisms and ocean global change. Nat. Microbiol. 2, 17058 (2017).40.Zakem, E. J. et al. Ecological control of nitrite in the upper ocean. Nat. Commun. 9, 1206 (2018).41.Axtell, R. L. Zipf distribution of U.S. firm sizes. Science 293, 1818–1820 (2001).42.Turchin, P. et al. Quantitative historical analysis uncovers a single dimension of complexity that structures global variation in human social organization. Proc. Natl Acad. Sci. USA 115, E144–E151 (2018).CAS 
    Article 

    Google Scholar 
    43.Press, W. H. Numerical Recipes in C (Cambridge Univ. Press, 1986). More

  • in

    Predicting species distributions and community composition using satellite remote sensing predictors

    1.Lewis, S. L. & Maslin, M. A. Defining the anthropocene. Nature 519, 171–180 (2015).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    2.Tilman, D. et al. Future threats to biodiversity and pathways to their prevention. Nature 546, 73–81 (2017).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    3.Pinto-Ledezma, J. N. & Rivero Mamani, M. L. Temporal patterns of deforestation and fragmentation in lowland Bolivia: Implications for climate change. Clim. Change 127, 43–54 (2014).ADS 
    Article 

    Google Scholar 
    4.Allen, J. M., Folk, R. A., Soltis, P. S., Soltis, D. E. & Guralnick, R. P. Biodiversity synthesis across the green branches of the tree of life. Nat. Plants 5, 11–13 (2019).PubMed 
    Article 

    Google Scholar 
    5.Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services, IPBES. Summary for policymakers of the global assessment report on biodiversity and ecosystem services. Accessed 15 Feb 2021. https://zenodo.org/record/3553579. https://doi.org/10.5281/ZENODO.3553579 (2019). 6.Cavender-Bares, J., Balvanera, P., King, E. & Polasky, S. Ecosystem service trade-offs across global contexts and scales. Ecol. Soc. 20, art22 (2015).Article 

    Google Scholar 
    7.Chaplin-Kramer, R. et al. Global modeling of nature’s contributions to people. Science 366, 255–258 (2019).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    8.Díaz, S. et al. Pervasive human-driven decline of life on Earth points to the need for transformative change. Science 366, eaax3100 (2019).Article 
    CAS 
    PubMed 

    Google Scholar 
    9.Watson, J. E. M. et al. Set a global target for ecosystems. Nature 578, 360–362 (2020).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    10.Jetz, W. et al. Essential biodiversity variables for mapping and monitoring species populations. Nat. Ecol. Evol. 3, 539–551 (2019).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    11.Mateo, R. G., Mokany, K. & Guisan, A. Biodiversity models: What if unsaturation is the rule?. Trends Ecol. Evol. 32, 556–566 (2017).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    12.Pereira, H. M. et al. Essential biodiversity variables. Science 339, 277–278 (2013).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    13.Ferrier, S. & Guisan, A. Spatial modelling of biodiversity at the community level. J. Appl. Ecol. 43, 393–404 (2006).Article 

    Google Scholar 
    14.Guisan, A. & Rahbek, C. SESAM—A new framework integrating macroecological and species distribution models for predicting spatio-temporal patterns of species assemblages: Predicting spatio-temporal patterns of species assemblages. J. Biogeogr. 38, 1433–1444 (2011).Article 

    Google Scholar 
    15.Cavender-Bares, J., Schweiger, A. K., Pinto-Ledezma, J. N. & Meireles, J. E. Applying remote sensing to biodiversity science. in Remote Sensing of Plant Biodiversity (eds. Cavender-Bares, J. et al.) 13–42 (Springer, 2020). https://doi.org/10.1007/978-3-030-33157-3_2.Chapter 

    Google Scholar 
    16.Fawcett, D. et al. Advancing retrievals of surface reflectance and vegetation indices over forest ecosystems by combining imaging spectroscopy, digital object models, and 3D canopy modelling. Remote Sens. Environ. 204, 583–595 (2018).ADS 
    Article 

    Google Scholar 
    17.Randin, C. F. et al. Monitoring biodiversity in the Anthropocene using remote sensing in species distribution models. Remote Sens. Environ. 239, 111626 (2020).ADS 
    Article 

    Google Scholar 
    18.Turner, W. Sensing biodiversity. Science 346, 301–302 (2014).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    19.D’Amen, M., Pradervand, J.-N. & Guisan, A. Predicting richness and composition in mountain insect communities at high resolution: A new test of the SESAM framework: Community-level models of insects. Glob. Ecol. Biogeogr. 24, 1443–1453 (2015).Article 

    Google Scholar 
    20.Pottier, J. et al. The accuracy of plant assemblage prediction from species distribution models varies along environmental gradients: Climate and species assembly predictions. Glob. Ecol. Biogeogr. 22, 52–63 (2013).Article 

    Google Scholar 
    21.Zurell, D. et al. Testing species assemblage predictions from stacked and joint species distribution models. J. Biogeogr. 47, 101–113 (2020).Article 

    Google Scholar 
    22.D’Amen, M. et al. Improving spatial predictions of taxonomic, functional and phylogenetic diversity. J. Ecol. 106, 76–86 (2018).Article 

    Google Scholar 
    23.Dobrowski, S. Z. et al. Modeling plant ranges over 75 years of climate change in California, USA: Temporal transferability and species traits. Ecol. Monogr. 81, 241–257 (2011).Article 

    Google Scholar 
    24.Soria-Auza, R. W. et al. Impact of the quality of climate models for modelling species occurrences in countries with poor climatic documentation: A case study from Bolivia. Ecol. Model. 221, 1221–1229 (2010).Article 

    Google Scholar 
    25.Rocchini, D. et al. Satellite remote sensing to monitor species diversity: Potential and pitfalls. Remote Sens. Ecol. Conserv. 2, 25–36 (2016).Article 

    Google Scholar 
    26.Schulte to Bühne, H. & Pettorelli, N. Better together: Integrating and fusing multispectral and radar satellite imagery to inform biodiversity monitoring, ecological research and conservation science. Methods Ecol. Evol. 9, 849–865 (2018).Article 

    Google Scholar 
    27.Hansen, M. C. et al. High-resolution global maps of 21st-century forest cover change. Science 342, 850–853 (2013).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    28.Hobi, M. L. et al. A comparison of Dynamic Habitat Indices derived from different MODIS products as predictors of avian species richness. Remote Sens. Environ. 195, 142–152 (2017).ADS 
    Article 

    Google Scholar 
    29.Radeloff, V. C. et al. The Dynamic Habitat Indices (DHIs) from MODIS and global biodiversity. Remote Sens. Environ. 222, 204–214 (2019).ADS 
    Article 

    Google Scholar 
    30.Pinto-Ledezma, J. N. & Cavender-Bares, J. Using remote sensing for modeling and monitoring species distributions. In Remote Sensing of Plant Biodiversity (eds. Cavender-Bares, J., Gamon, J. A. & Townsend, P. A.) 199–223 (Springer International Publishing, 2020). https://doi.org/10.1007/978-3-030-33157-3_9.31.Fernández, N., Ferrier, S., Navarro, L. M. & Pereira, H. M. Essential biodiversity variables: Integrating in-situ observations and remote sensing through modeling. In Remote Sensing of Plant Biodiversity (eds. Cavender-Bares, J., Gamon, J. A. & Townsend, P. A.) 485–501 (Springer International Publishing, 2020). https://doi.org/10.1007/978-3-030-33157-3_18.32.Skidmore, A. K. et al. Priority list of biodiversity metrics to observe from space. Nat. Ecol. Evol. https://doi.org/10.1038/s41559-021-01451-x (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    33.Saatchi, S., Buermann, W., ter Steege, H., Mori, S. & Smith, T. B. Modeling distribution of Amazonian tree species and diversity using remote sensing measurements. Remote Sens. Environ. 112, 2000–2017 (2008).ADS 
    Article 

    Google Scholar 
    34.He, K. S. et al. Will remote sensing shape the next generation of species distribution models?. Remote Sens. Ecol. Conserv. 1, 4–18 (2015).Article 

    Google Scholar 
    35.Cord, A. F., Meentemeyer, R. K., Leitão, P. J. & Václavík, T. Modelling species distributions with remote sensing data: Bridging disciplinary perspectives. J. Biogeogr. 40, 2226–2227 (2013).Article 

    Google Scholar 
    36.Jetz, W. et al. Monitoring plant functional diversity from space. Nat. Plants 2, 16024 (2016).PubMed 
    Article 

    Google Scholar 
    37.Scherrer, D., D’Amen, M., Fernandes, R. F., Mateo, R. G. & Guisan, A. How to best threshold and validate stacked species assemblages? Community optimisation might hold the answer. Methods Ecol. Evol. 9, 2155–2166 (2018).Article 

    Google Scholar 
    38.Cavender-Bares, J. Diversification, adaptation, and community assembly of the American oaks (Quercus), a model clade for integrating ecology and evolution. New Phytol. 221, 669–692 (2019).PubMed 
    Article 

    Google Scholar 
    39.Cavender-Bares, J., Ackerly, D. D., Baum, D. A. & Bazzaz, F. A. Phylogenetic overdispersion in Floridian oak communities. Am. Nat. 163, 823–843 (2004).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    40.Cavender-Bares, J. et al. The role of diversification in community assembly of the oaks (Quercus L.) across the continental U.S. Am. J. Bot. 105, 565–586 (2018).PubMed 
    Article 

    Google Scholar 
    41.Colwell, R. K. & Rangel, T. F. Hutchinson’s duality: The once and future niche. Proc. Natl. Acad. Sci. 106, 19651–19658 (2009).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    42.Townsend Peterson, A. et al. Ecological Niches and Geographic Distributions. (Princeton University Press, 2011). Book 

    Google Scholar 
    43.Cavender-Bares, J., Fontes, G. C. & Pinto-Ledezma, J. Open questions in understanding the adaptive significance of plant functional trait variation within a single lineage. New Phytol. https://doi.org/10.1111/nph.16652 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    44.Cavender-Bares, J., Kitajima, K. & Bazzaz, F. A. Multiple trait associations in relation to habitat differentiation among 17 Floridian oak species. Ecol. Monogr. 74, 635–662 (2004).Article 

    Google Scholar 
    45.Menges, E. S. & Hawkes, C. V. Interactive effects of fire and microhabitat on plants of Florida scrub. Ecol. Appl. 8, 935–946 (1998).Article 

    Google Scholar 
    46.Calabrese, J. M., Certain, G., Kraan, C. & Dormann, C. F. Stacking species distribution models and adjusting bias by linking them to macroecological models: Stacking species distribution models. Glob. Ecol. Biogeogr. 23, 99–112 (2014).Article 

    Google Scholar 
    47.Hurlbert, A. H. & Jetz, W. Species richness, hotspots, and the scale dependence of range maps in ecology and conservation. Proc. Natl. Acad. Sci. 104, 13384–13389 (2007).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    48.Pinto-Ledezma, J. N., Jahn, A. E., Cueto, V. R., Diniz-Filho, J. A. F. & Villalobos, F. Drivers of phylogenetic assemblage structure of the Furnariides, a widespread clade of lowland neotropical birds. Am. Nat. 193, E41–E56 (2019).PubMed 
    Article 

    Google Scholar 
    49.Gamon, J. A. et al. Consideration of scale in remote sensing of biodiversity. In Remote Sensing of Plant Biodiversity (eds. Cavender-Bares, J., Gamon, J. A. & Townsend, P. A.) 425–447 (Springer International Publishing, 2020). https://doi.org/10.1007/978-3-030-33157-3_16.50.Pollock, L. J. et al. Understanding co-occurrence by modelling species simultaneously with a Joint Species Distribution Model (JSDM). Methods Ecol. Evol. 5, 397–406 (2014).Article 

    Google Scholar 
    51.Ovaskainen, O. et al. How to make more out of community data? A conceptual framework and its implementation as models and software. Ecol. Lett. 20, 561–576 (2017).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    52.Ovaskainen, O. Joint Species Distribution Modelling: with Applications in R (Cambridge University Press, 2020).Book 

    Google Scholar 
    53.Poggiato, G. et al. On the interpretations of joint modeling in community ecology. Trends Ecol. Evol. https://doi.org/10.1016/j.tree.2021.01.002 (2021).Article 
    PubMed 

    Google Scholar 
    54.Wilkinson, D. P., Golding, N., Guillera-Arroita, G., Tingley, R. & McCarthy, M. A. Defining and evaluating predictions of joint species distribution models. Methods Ecol. Evol. 12, 394–404 (2021).Article 

    Google Scholar 
    55.Bystrova, D. et al. Clustering species with residual covariance matrix in joint species distribution models. Front. Ecol. Evol. 9, 601384 (2021).Article 

    Google Scholar 
    56.Mateo, R. G. et al. Hierarchical species distribution models in support of vegetation conservation at the landscape scale. J. Veg. Sci. 30, 386–396 (2019).ADS 
    Article 

    Google Scholar 
    57.Petitpierre, B. et al. Will climate change increase the risk of plant invasions into mountains?. Ecol. Appl. 26, 530–544 (2016).PubMed 
    Article 

    Google Scholar 
    58.Cavender-Bares, J. et al. Harnessing plant spectra to integrate the biodiversity sciences across biological and spatial scales. Am. J. Bot. 104, 966–969 (2017).PubMed 
    Article 

    Google Scholar 
    59.Schweiger, A. K. et al. Spectral Niches Reveal Taxonomic Identity and Complementarity in Plant Communities. (2020) https://doi.org/10.1101/2020.04.24.060483. 60.Cavender-Bares, J. et al. Associations of leaf spectra with genetic and phylogenetic variation in oaks: Prospects for remote detection of biodiversity. Remote Sens. 8, 221 (2016).ADS 
    Article 

    Google Scholar 
    61.Meireles, J. E. et al. Leaf reflectance spectra capture the evolutionary history of seed plants. New Phytol. 228, 485–493 (2020).PubMed 
    PubMed Central 
    Article 

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

    Google Scholar 
    63.Alonso, K. et al. Data products, quality and validation of the DLR earth sensing imaging spectrometer (DESIS). Sensors 19, 4471 (2019).ADS 
    CAS 
    PubMed Central 
    Article 

    Google Scholar 
    64.Stavros, E. N. et al. ISS observations offer insights into plant function. Nat. Ecol. Evol. 1, 0194 (2017).Article 

    Google Scholar 
    65.Féret, J.-B. & Asner, G. P. Mapping tropical forest canopy diversity using high-fidelity imaging spectroscopy. Ecol. Appl. 24, 1289–1296 (2014).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    66.Cavender-Bares, J. et al. BII-Implementation: The causes and consequences of plant biodiversity across scales in a rapidly changing world. Res. Ideas Outcomes 7, e63850 (2021).Article 

    Google Scholar 
    67.Hipp, A. L. et al. Sympatric parallel diversification of major oak clades in the Americas and the origins of Mexican species diversity. New Phytol. 217, 439–452 (2018).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    68.Cavender-Bares, J. et al. Phylogeny and biogeography of the American live oaks (Quercus subsection Virentes): A genomic and population genetics approach. Mol. Ecol. 24, 3668–3687 (2015).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    69.Aiello-Lammens, M. E., Boria, R. A., Radosavljevic, A., Vilela, B. & Anderson, R. P. spThin: An R package for spatial thinning of species occurrence records for use in ecological niche models. Ecography 38, 541–545 (2015).Article 

    Google Scholar 
    70.Lobo, J. M., Jiménez-Valverde, A. & Hortal, J. The uncertain nature of absences and their importance in species distribution modelling. Ecography 33, 103–114 (2010).Article 

    Google Scholar 
    71.Barnett, D. T. et al. The plant diversity sampling design for The National Ecological Observatory Network. Ecosphere 10, e02603 (2019).
    Google Scholar 
    72.Deblauwe, V. et al. Remotely sensed temperature and precipitation data improve species distribution modelling in the tropics: Remotely sensed climate data for tropical species distribution models. Glob. Ecol. Biogeogr. 25, 443–454 (2016).Article 

    Google Scholar 
    73.Hijmans, R. J., Phillips, S., Leathwick, J. & Elith, J. dismo: Species Distribution Modeling. R package version 1.3. https://CRAN.R-project.org/package=dismo (2020).74.Myneni, R. B. et al. Global products of vegetation leaf area and fraction absorbed PAR from year one of MODIS data. Remote Sens. Environ. 83, 214–231 (2002).ADS 
    Article 

    Google Scholar 
    75.Gower, S. T., Kucharik, C. J. & Norman, J. M. Direct and indirect estimation of leaf area index, fAPAR, and net primary production of terrestrial ecosystems. Remote Sens. Environ. 70, 29–51 (1999).ADS 
    Article 

    Google Scholar 
    76.Reich, P. B. Key canopy traits drive forest productivity. Proc. R. Soc. B Biol. Sci. 279, 2128–2134 (2012).Article 

    Google Scholar 
    77.Xiao, Z. et al. Use of general regression neural networks for generating the GLASS leaf area index product from time-series MODIS surface reflectance. IEEE Trans. Geosci. Remote Sens. 52, 209–223 (2014).ADS 
    Article 

    Google Scholar 
    78.Soberón, J. Grinnellian and Eltonian niches and geographic distributions of species. Ecol. Lett. 10, 1115–1123 (2007).PubMed 
    Article 

    Google Scholar 
    79.Farr, T. G. et al. The shuttle radar topography mission. Rev. Geophys. 45, RG2004 (2007).ADS 
    Article 

    Google Scholar 
    80.Dubuis, A. et al. Predicting spatial patterns of plant species richness: A comparison of direct macroecological and species stacking modelling approaches: Predicting plant species richness. Divers. Distrib. 17, 1122–1131 (2011).Article 

    Google Scholar 
    81.Schoener, T. W. Anolis lizards of Bimini: Resource partition in a complex fauna. Ecology 49, 704–726 (1968).Article 

    Google Scholar 
    82.Barve, N. et al. The crucial role of the accessible area in ecological niche modeling and species distribution modeling. Ecol. Model. 222, 1810–1819 (2011).Article 

    Google Scholar 
    83.Cooper, J. C. & Soberón, J. Creating individual accessible area hypotheses improves stacked species distribution model performance. Glob. Ecol. Biogeogr. 27, 156–165 (2018).Article 

    Google Scholar 
    84.Barbet-Massin, M., Jiguet, F., Albert, C. H. & Thuiller, W. Selecting pseudo-absences for species distribution models: How, where and how many? How to use pseudo-absences in niche modelling?. Methods Ecol. Evol. 3, 327–338 (2012).Article 

    Google Scholar 
    85.Carlson, C. J. et al. The global distribution of Bacillus anthracis and associated anthrax risk to humans, livestock and wildlife. Nat. Microbiol. 4, 1337–1343 (2019).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    86.Chipman, H. A., George, E. I. & McCulloch, R. E. BART: Bayesian additive regression trees. Ann. Appl. Stat. 4, 266–298 (2010).MathSciNet 
    MATH 
    Article 

    Google Scholar 
    87.Yen, J. D. L., Thomson, J. R., Vesk, P. A. & Mac Nally, R. To what are woodland birds responding? Inference on relative importance of in-site habitat variables using several ensemble habitat modelling techniques. Ecography 34, 946–954 (2011).Article 

    Google Scholar 
    88.Carlson, C. J. embarcadero: Species distribution modelling with Bayesian additive regression trees in r. Methods Ecol. Evol. 11, 850–858 (2020).Article 

    Google Scholar 
    89.Dorie, V. dbarts: Discrete Bayesian Additive Regression Trees Sampler. (2020).90.Hastie, T. & Tibshirani, R. Bayesian backfitting. Stat. Sci. 15(3), 196–223 (2000). MathSciNet 
    MATH 
    Article 

    Google Scholar 
    91.Allouche, O., Tsoar, A. & Kadmon, R. Assessing the accuracy of species distribution models: Prevalence, kappa and the true skill statistic (TSS): Assessing the accuracy of distribution models. J. Appl. Ecol. 43, 1223–1232 (2006).Article 

    Google Scholar 
    92.Di Cola, V. et al. ecospat: An R package to support spatial analyses and modeling of species niches and distributions. Ecography 40, 774–787 (2017).Article 

    Google Scholar 
    93.Hipp, A. L. et al. Genomic landscape of the global oak phylogeny. New Phytol. 226, 1198–1212 (2020).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    94.Webb, C. O., Ackerly, D. D., McPeek, M. A. & Donoghue, M. J. Phylogenies and community ecology. Annu. Rev. Ecol. Syst. 33, 475–505 (2002).Article 

    Google Scholar 
    95.Kembel, S. W. et al. Picante: R tools for integrating phylogenies and ecology. Bioinformatics 26, 1463–1464 (2010).CAS 
    PubMed 
    Article 

    Google Scholar 
    96.Kruschke, J. Doing Bayesian Data Analysis, 2nd Ed. (2014).97.Mills, J. A. & Parent, O. Bayesian MCMC estimation. In Handbook of Regional Science (eds Fischer, M. M. & Nijkamp, P.) 1571–1595 (Springer, Berlin, 2014). https://doi.org/10.1007/978-3-642-23430-9_89.Chapter 

    Google Scholar 
    98.Carpenter, B. et al. Stan : A probabilistic programming language. J. Stat. Softw. 76, 1–32 (2017).Article 

    Google Scholar 
    99.Goodrich, B., Gabry, J., Ali, I. & Brilleman, S. rstanarm: Bayesian applied regression modeling via Stan. (2020).100.Bürkner, P.-C. brms: An R package for Bayesian multilevel models using Stan. J. Stat. Softw. 80, 1–28 (2017).Article 

    Google Scholar 
    101.Makowski, D., Ben-Shachar, M. & Lüdecke, D. bayestestR: Describing effects and their uncertainty, existence and significance within the Bayesian framework. J. Open Source Softw. 4, 1541 (2019).ADS 
    Article 

    Google Scholar  More

  • in

    An approach to assess stress in response to drive hunts using cortisol levels of wild boar (Sus scrofa)

    1.Palme, R. Monitoring stress hormone metabolites as a useful, non-invasive tool for welfare assessment in farm animals. Anim. Welf. 21, 331–337 (2012).CAS 
    Article 

    Google Scholar 
    2.Jankord, R. & Herman, J. P. Limbic regulation of hypothalamo-pituitary-adrenocortical function during acute and chronic stress. Ann. N. Y. Acad. Sci. 1148, 64–73 (2008).ADS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    3.Romero, L. M. Physiological stress in ecology: Lessons from biomedical research. Trends Ecol. Evol. 19, 249–255 (2004).PubMed 
    Article 

    Google Scholar 
    4.Haase, C. G., Long, A. K. & Gillooly, J. F. Energetics of stress: Linking plasma cortisol levels to metabolic rate in mammals. Biol. Lett. 12, 20150867 (2016).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    5.Selye, H. A syndrome produced by diverse nocuous agents. Nature 1936, 32 (1936).ADS 
    Article 

    Google Scholar 
    6.Fink, G. Stress Science: Neuroendocrinology (Academic Press, Elsevier Science, 2010).
    Google Scholar 
    7.Hing, S., Narayan, E. J., Thompson, R. C. A. & Godfrey, S. S. The relationship between physiological stress and wildlife disease: Consequences for health and conservation. Wildl. Res. 43, 51 (2016).Article 

    Google Scholar 
    8.Tsigos, C. & Chrousos, G. P. Hypothalamic-pituitary-adrenal axis, neuroendocrine factors and stress. J. Psychosom. Res. 53, 865–871 (2002).PubMed 
    Article 

    Google Scholar 
    9.Tryphonopoulos, P. D., Letourneau, N. & Azar, R. Approaches to salivary cortisol collection and analysis in infants. Biol. Res. Nurs. 16, 398–408 (2014).PubMed 
    Article 

    Google Scholar 
    10.Palme, R. Non-invasive measurement of glucocorticoids: Advances and problems. Physiol. Behav. 199, 229–243 (2019).CAS 
    PubMed 
    Article 

    Google Scholar 
    11.Russell, E., Koren, G., Rieder, M. & Van Uum, S. Hair cortisol as a biological marker of chronic stress: Current status, future directions and unanswered questions. Psychoneuroendocrinology 37, 589–601 (2012).CAS 
    PubMed 
    Article 

    Google Scholar 
    12.McEwen, B. S. Central effects of stress hormones in health and disease: Understanding the protective and damaging effects of stress and stress mediators. Eur. J. Pharmacol. 583, 174–185 (2008).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    13.O’Connor, E. A. et al. The impact of chronic environmental stressors on growing pigs, Sus scrofa (Part 1): Stress physiology, production and play behaviour. Animal 4, 1899–1909 (2010).PubMed 
    Article 

    Google Scholar 
    14.Kadarmideen, H. N. & Janss, L. L. G. Population and systems genetics analyses of cortisol in pigs divergently selected for stress. Physiol. Genomics 29, 57–65 (2007).CAS 
    PubMed 
    Article 

    Google Scholar 
    15.Romano, M. C. et al. Stress in wildlife species: Noninvasive monitoring of glucocorticoids. NeuroImmunoModulation 17, 209–212 (2010).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    16.Sales, L. P. et al. Niche conservatism and the invasive potential of the wild boar. J. Anim. Ecol. 86, 1214–1223 (2017).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    17.Briedermann, L. Schwarzwild (Franckh-Kosmos Verlags-GmnH & Co. KG, 2009).
    Google Scholar 
    18.Keuling, O. et al. Eurasian wild boar Sus scrofa (Linnaeus, 1758). In Ecology, Conservation and Management of Wild Pigs and Peccaries (eds Melletti, M. & Meijaard, E.) 202–233 (Cambridge University Press, 2018).
    Google Scholar 
    19.Niedersächsisches Ministerium für Ernährung, Landwirtschaft und Verbraucherschutz. Aktuelle Jagdzeiten in Niedersachsen (konsolidierte Fassung) Stand: 25. Januar 2021 inkl. Verordnung zur Durchführung des Nieders. Jagdgesetzes (DVO-NJagdG) vom 23. Mai 2008 (Nds. GVBl. S. 194), zuletzt geändert durch Verordnung vom 18. Januar 2021 (Nds. GVBl. S. 24). (2021). https://www.ml.niedersachsen.de/download/163729/Aktuelle_Jagdzeiten_in_Niedersachsen_Stand_25.01.2021_nicht_vollstaendig_barrierefrei_.pdf. Accessed 01 June 2021.20.Casas-Díaz, E. et al. Hematologic and biochemical reference intervals for Wild Boar (Sus scrofa) captured by cage trap. Vet. Clin. Pathol. 44, 215–222 (2015).PubMed 
    Article 

    Google Scholar 
    21.Gentsch, R. P., Kjellander, P. & Röken, B. O. Cortisol response of wild ungulates to trauma situations: Hunting is not necessarily the worst stressor. Eur. J. Wildl. Res. 64, 11 (2018).Article 

    Google Scholar 
    22.Adcock, S. J. J., Martin, G. M. & Walsh, C. J. The stress response and exploratory behaviour in Yucatan minipigs (Sus scrofa): Relations to sex and social rank. Physiol. Behav. 152, 194–202 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    23.Bratton, S. P. The effect of the European wild boar (Sus scrofa) on gray beech forest in the great smokey mountains. Ecology 56, 1356–1366 (1975).Article 

    Google Scholar 
    24.Singer, F. J., Swank, W. T. & Clebsh, E. E. C. The effects of wild pig rooting in a deciduous forest. J. Wildl. Manage. 48, 464–473 (1984).CAS 
    Article 

    Google Scholar 
    25.Wlazelko, M. & Labudzki, L. Über Nahrungskomponenten und trophische Stellung des Schwarzwildes im Forschungsgebiet Zielonka. Z. Jagdwiss. 38, 81–87 (1992).
    Google Scholar 
    26.Killian, G., Miller, L., Rhyan, J. & Doten, H. Immunocontraception of Florida feral swine with a single-dose GnRH vaccine. Am. J. Reprod. Immunol. 55, 378–384 (2006).CAS 
    PubMed 
    Article 

    Google Scholar 
    27.Gortázar, C., Ferroglio, E., Höfle, U., Frölich, K. & Vicente, J. Diseases shared between wildlife and livestock: A European perspective. Eur. J. Wildl. Res. 53, 241–256 (2007).Article 

    Google Scholar 
    28.Gräber, R., Strauß, E. & Johanshon, S. Wild und Jagd—Landesjagdbericht 2017/2018 (Niedersächsisches Ministerium für Ernährung, Landwirtschaft und Verbraucherschutz, Hannover, 2018).29.Wölfel, H. Bewegungsjagden (Leopold Stocker Verlag, 2003).30.Eisenbarth, E. & Ophoven, E. Bewegungsjagd auf Schalenwild (Franckh-Kosmos Verlags-GmbH & Co., 2002).
    Google Scholar 
    31.Böhm, E. Drückjagd auf Sauen (Neumann-Neudamm, 2004).
    Google Scholar 
    32.Bradshaw, E. L. & Bateson, P. Welfare implications of culling red deer (Cervus elaphus). Anim. Welf. 9, 3–24 (2000).
    Google Scholar 
    33.Sheriff, M. J., Dantzer, B., Delehanty, B., Palme, R. & Boonstra, R. Measuring stress in wildlife: Techniques for quantifying glucocorticoids. Oecologia 166, 869–887 (2011).ADS 
    PubMed 
    Article 

    Google Scholar 
    34.Hellhammer, D. H., Wüst, S. & Kudielka, B. M. Salivary cortisol as a biomarker in stress research. Psychoneuroendocrinology 34, 163–171 (2009).CAS 
    PubMed 
    Article 

    Google Scholar 
    35.Palme, R., Rettenbacher, S., Touma, C., El-Bahr, S. M. & Möstl, E. Stress hormones in mammals and birds: Comparative aspects regarding metabolism, excretion, and noninvasive measurement in fecal samples. Ann. N. Y. Acad. Sci. 1040, 162–171 (2005).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    36.Kanitz, E., Otten, W., Tuchscherer, M. & Manteuffel, G. Effects of prenatal stress on corticosteroid receptors and monoamine concentrations in limbic areas of suckling piglets (Sus scrofa) at different ages. J. Vet. Med. Ser. A 50, 132–139 (2003).CAS 
    Article 

    Google Scholar 
    37.Campbell, E. A. et al. Plasma corticotropin-releasing hormone concentrations during pregnancy and parturition. J. Clin. Endocrinol. Metab. 64, 1054–1059 (1987).CAS 
    PubMed 
    Article 

    Google Scholar 
    38.Seth, S., Lewis, A. J. & Galbally, M. Perinatal maternal depression and cortisol function in pregnancy and the postpartum period: A systematic literature review. BMC Pregn. Childbirth 16, 124 (2016).Article 
    CAS 

    Google Scholar 
    39.Gethöffer, F. Reproduktionsparameter und Saisonalität der Fortpflanzung des Wildschweins (Sus scrofa) in drei Untersuchungsgebieten Deutschlands (University of Veterinary Medicine Hannover, 2005).
    Google Scholar 
    40.Frauendorf, M., Gethöffer, F., Siebert, U. & Keuling, O. The influence of environmental and physiological factors on the litter size of wild boar (Sus scrofa) in an agriculture dominated area in Germany. Sci. Total Environ. 541, 877–882 (2016).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    41.Gethöffer, F., Sodeikat, G. & Pohlmeyer, K. Reproductive parameters of wild boar (Sus scrofa) in three different parts of Germany. Eur. J. Wildl. Res. 53, 287–297 (2007).Article 

    Google Scholar 
    42.DWD. Deutscher Wetterdienst—Wetter und Klima—Klimadaten (2019). https://www.dwd.de. Accessed 01 Oct 2019.43.Keuling, O., Stier, N. & Roth, M. Annual and seasonal space use of different age classes of female wild boar Sus scrofa L.. Eur. J. Wildl. Res. 54, 403–412 (2008).Article 

    Google Scholar 
    44.Malmsten, A., Jansson, G., Lundeheim, N. & Dalin, A.-M. The reproductive pattern and potential of free ranging female wild boars (Sus scrofa) in Sweden. Acta Vet. Scand. 59, 52 (2017).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    45.R Core Team. R: A Language and Environment for Statistical Computing Version R3.5.2.  R Foundation for Statistical Computing, Vienna, Austria.  http://www.R-project.org/ (2018).46.Dunn, O. J. Multiple comparisons using rank sums. Technometrics 6, 241–252 (1964).Article 

    Google Scholar 
    47.Ogle, D. H., Wheeler, P. & Dinno, A. FSA: Fisheries Stock Analysis. R Package Version 0.8.25. https://github.com/droglenc/FSA (2019).48.Kassambara, A. ggpubr: ‘ggplot2’ Based Publication Ready Plots. R Package Version 0.2.3. http://www.sthda.com/english/rpkgs/ggpubr (2019).49.Palme, R. Measuring fecal steroids: Guidelines for practical application. Ann. N. Y. Acad. Sci. 1046, 75–80 (2005).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    50.Cockrem, J. F. Individual variation in glucocorticoid stress responses in animals. Gen. Comp. Endocrinol. 181, 45–58 (2013).CAS 
    PubMed 
    Article 

    Google Scholar 
    51.Mormède, P. et al. Exploration of the hypothalamic-pituitary-adrenal function as a tool to evaluate animal welfare. Physiol. Behav. 92, 317–339 (2007).PubMed 
    Article 
    CAS 

    Google Scholar 
    52.Goymann, W. Noninvasive monitoring of hormones in bird droppings: Physiological validation, sampling, extraction, sex differences, and the influence of diet on hormone metabolite levels. Ann. N. Y. Acad. Sci. 1046, 35–53 (2005).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    53.Guilliams, T. G. & Edwards, L. Chronic stress and the HPA axis: Clinical assessment and therapeutic considerations. Stand. 9, 1–12 (2010).
    Google Scholar 
    54.Merta, D., Mocala, P., Pomykacz, M. & Frackowiak, W. Autumn-winter diet and fat reserves of wild boars (Sus scrofa) inhabiting forest and forest-farmland environment in south-western Poland. Folia Zool. 63, 95–102 (2014).Article 

    Google Scholar 
    55.Poteaux, C. et al. Socio-genetic structure and mating system of a wild boar population. J. Zool. 278, 116–125 (2009).Article 

    Google Scholar 
    56.Kaminski, G., Brandt, S., Baubet, E. & Baudoin, C. Life-history patterns in female wild boars (Sus scrofa): Mother–daughter postweaning associations. Can. J. Zool. 83, 474–480 (2005).Article 

    Google Scholar 
    57.Krause, J. & Ruxton, G. D. Living in Groups. Oxford Series in Ecology and Evolution (Oxford University Press, 2002).
    Google Scholar 
    58.Kudielka, B. M. & Kirschbaum, C. Sex differences in HPA axis responses to stress: A review. Biol. Psychol. 69, 113–132 (2005).PubMed 
    Article 

    Google Scholar 
    59.Balhara, Y. S., Verma, R. & Gupta, C. Gender differences in stress response: Role of developmental and biological determinants. Ind. Psychiatry J. 20, 4 (2012).Article 

    Google Scholar 
    60.Sutherland, M. A., Rodriguez-Zas, S. L., Ellis, M. & Salak-Johnson, J. L. Breed and age affect baseline immune traits, cortisol, and performance in growing pigs. J. Anim. Sci. 83, 2087–2095 (2005).CAS 
    PubMed 
    Article 

    Google Scholar 
    61.Foury, A. et al. Stress hormones, carcass composition and meat quality in Large White × Duroc pigs. Meat Sci. 69, 703–707 (2005).CAS 
    PubMed 
    Article 

    Google Scholar 
    62.Ruis, M. A. W. et al. The circadian rhythm of salivary cortisol in growing pigs: Effects of age, gender, and stress. Physiol. Behav. 62, 623–630 (1997).CAS 
    PubMed 
    Article 

    Google Scholar  More

  • in

    Environmental risk evaluation of overseas mining investment based on game theory and an extension matter element model

    Data sourcesThe data come from the Ministry of Commerce of the People’s Republic of China’s 2019 Guide to Foreign Investment and Cooperation Country, as well as the websites and research literature from the Fraser Institute and the World Bank. The datasets include 14 factors that influence the environmental risk of overseas mining investment in the Philippines are summarized in Tables 1 and 2. The specific reasons that we choose these data in the Philippines are as follows:Table 1 Evaluation factors and grading standards of environmental risk for an evaluation of overseas mining investment in the Philippines.Full size tableTable 2 Risk index data.Full size tableTable 3 Correlation function value of each evaluation index used in an evaluation of overseas mining investment in the Philippines.Full size tableThe Philippines is a multi-ethnic island nation in Southeast Asia located in the western Pacific Ocean. The country has a total land area of 299,700 km2 and a population of 101 million. The Philippines is rich in mineral resources, and the area of known mineralization accounts for 30% of the land area in the country. According to the National Bureau of Geology and Mining in the Philippines, gold, copper, nickel, and chromium reserves rank third, fourth, fifth, and sixth in the world, respectively, in terms of mineral reserves per unit area. Nonferrous metal mining in the Philippines has great potential. To date, 13 types of metal minerals have been discovered, including gold, copper, nickel, aluminium, chromium, silver, lead, and zinc, with total reserves of 7.1 billion tons. Twenty-nine types of nonmetallic minerals have also been discovered with total reserves of 51 billion tons. The Philippines is an important producer and exporter of metallic mineral resources such as copper and nickel6,33.The Philippines has been one of the countries most in favour of overseas mining investment in the region near China. Before the mid-1990s, the Philippines was a favoured country for international mining investors; however, in the late 1990s, changes to national policies and social unrest led to a decline in the mining investment environment. Since January 2003, President Arroyo has proposed a reform of the mining development strategy in the Philippines, and the mining investment environment has improved. However, combined with the political, religious and security issues in the Philippines, especially the peoples’ attitude towards foreign investment, the current mining policy environment in the Philippines is not ideal. Therefore, to comprehensively and objectively understand and analyse the mining investment environment in the Philippines, relevant documents were collated and analysed. Following the principles of importance, practicality, scientificity and systematicness in the design of the index system, the accepted classification rules and data released by authoritative agencies such as the World Bank were used for the evaluation basis1,2,3,4,5,6,7,8, which selected 14 factors that have bearing on political policy, economic, financial, sociocultural, and infrastructure risks. The classification standard and valuation of each index are provided in Table 1. According to the classification standard and valuation index objectives, the risks were divided into five levels (i.e., I–V, which reflect high, higher, general, lower, and low risks, respectively). The Philippines’ risk index data are listed in Table 2.Determination of index weights: analytical hierarchy processThis method integrates quantitative and qualitative evaluations to improve the accuracy of decision making32,33,34,35,36,37,38. The basic principles and steps of the AHP method are as follows:Step 1: The complex problem is decomposed to make it multi-element in nature.Step 2: These elements are grouped, and a hierarchical structural model is established.Step 3: A discrimination matrix is constructed, and any two factors are compared with a 1–9 scaling method to obtain the relative importance of each index at each level, which can be expressed quantitatively.Step 4: The largest eigenvalue and the corresponding eigenvector of the discrimination matrix are calculated using the mathematical method, where the eigenvectors and weight coefficient values are listed in terms of the importance of the evaluation factors.Step 5: The consistency of the discrimination matrix is tested based on the consistency index ( CI) calculated as (CI = frac{{{lambda_{max }} – n}}{n – 1}) as well as with the average random consistency index (RI). If the random consistency ratio (CR = frac{CI}{{RI}} < 0.10), then the results of the hierarchy analysis are considered to be consistent, and the resulting weight distribution values are reasonable. If this is not the case, then the weight coefficient values should be redistributed to adjust the values.Entropy weight theoryIn information theory, the importance of studying the degree of dispersion of the whole system is central to the entropy method. The specific steps for these calculations are as follows:Step 1: Data collection and sorting: The initial evaluation matrix composed of (m) evaluation indexes and (n) evaluation objects is as follows:$$ {{text{X}}_{{text{ij}}}} = left[ {begin{array}{*{20}{c}} {{x_{11}}}&{{x_{12}}}& cdots &{{x_{1{text{n}}}}} \ {{x_{21}}}&{{x_{22}}}& cdots &{{x_{2n}}} \ vdots & vdots & vdots & vdots \ {{x_{m1}}}&{{x_{m2}}}& cdots &{{x_{mn}}} end{array}} right] $$ (1) Step 2: Data standardization: All index values ( {x_{ij}}) in matrix ( {X_{ij}}) are normalized as follows:$$ {text{x}}_{ij}^{prime} = {raise0.7exhbox{${{x_{ij}}}$} !mathord{left/ {vphantom {{{x_{ij}}} {sumlimits_{i = 1}^m {{x_{ij}}} }}}right.kern-nulldelimiterspace}!lower0.7exhbox{${sumlimits_{i = 1}^m {{x_{ij}}} }$}} $$ (2) Step 3: Calculation of information entropy: The entropy of each evaluation index can be obtained from$$ {E_i} = frac{{sumlimits_{j = 1}^n {x_{ij}^{prime}ln x_{ij}^{prime}} }}{ln n} $$ (3) Step 4: Calculation weight: The weight of each evaluation index can be calculated as follows:$$ {w_i} = frac{{1 - {E_i}}}{{sumlimits_{i = 1}^m {left( {1 - {E_i}} right)} }} $$ (4) where ( {w_j}) is the index weight and ( sumlimits_{j = 1}^n {{w_j} = 1} ). The larger the entropy weight is, the greater the effect of the index on the scheme, in that it contains and transmits more decision information that has a greater influence on the final evaluation decision39,40,41,42,43,44.Combination weighting model based on game theoryThis approach differs from the traditional simple linear combination weighting method. The central idea of this approach is to “coordinate conflicts and maximize benefits” by comprehensively considering the relationship between the indexes, balancing the subjective and objective weights, and optimising the index weight values. The basic algorithm is as follows:Construction of the basic weight vector setAssuming that ( H) weight values are obtained using the ( H) weighting method, the basic weight vector set of the ( H) method is$$ {w_k} = left( {{w_{k1}},{w_{k2}}, cdots {w_{kn}}} right),k = 1,2, cdots ,H $$ (5) Any linear combination of ( H) weight vectors is$$ w = sumlimits_{k = 1}^H {{a_k}{w_k}^T} ,{a_k} > 0 $$
    (6)
    where ( {a_k}) is the linear combination coefficient, and ( w) is the comprehensive index weight value of the ( H) weight set.Optimal combination weightTo find the balance between the different weights, the optimal effect weight vector ( W) was obtained. In the calculation process, it is converted into an optimisation of the weight coefficient ( {a_k}) to minimise the deviation between ( w) and ( {w_k}), as follows:$$ minleft| {sumlimits_{j = 1}^H {{a_j}{W_j}^T – {W_i}^T} } right|,i = 1,2, cdots ,H;j = 1,2 cdots ,H $$
    (7)
    From the differential properties of the matrix, the first-order derivative condition for the optimisation of Eq. (7) becomes$$ sumlimits_{j = 1}^H {a_j} {W_i}W_j^T = {W_i}W_i^T $$
    (8)
    By solving Eq. (8), the combination coefficients ( left[ {{a_1},{a_2}, cdots ,{a_H}} right]) can be obtained and normalised according to ( a_k^* = {a_k}/sumlimits_{k = 1}^H {a_k} ). The final combination index weight is ( W = sumlimits_{k = 1}^H {a_k^*W_k^T} ,k = 1,2, cdots ,H) 31,32.Workflow of extension matter element theoryThe theoretical basis of extenics involves the matter element and extension set theories, and its logical cell is the matter element. As such, extenics introduces the concept of the matter element that organically combines quality and quantity. It is a triple group composed of things, features, and quantity values for things, which are depicted as R = (things, features, quantity values). The matter element concept correctly describes the relationship between quality and quantity, and it can be more appropriate to describe the change process of objective things. Different objects can have the same characteristic element and are represented by the matter element with the same characteristics. For convenience, many matter elements with the same characteristics are expressed in a simple way.Determination of the classical and joint domains$$ {R_{ij}} = left( {{N_j},{C_i},{V_{ij}}} right) = left[ {begin{array}{*{20}{c}} {N_j}&{C_1}&{{V_{1j}}} \ {}&{C_2}&{{V_{2j}}} \ {}& vdots & vdots \ {}&{C_i}&{{V_{ij}}} end{array}} right] = left[ {begin{array}{*{20}{c}} {N_j}&{C_1}&{left( {{a_{1j}},{b_{1j}}} right)} \ {}&{C_2}&{left( {{a_{2j}},{b_{2j}}} right)} \ {}& vdots & vdots \ {}&{C_i}&{({a_{ij}},{b_{ij}})} end{array}} right] $$
    (9)

    Equation (9) is a matter element body with the same characteristics of a matter element with the same characteristics ( {R_{ij}}), in which ( {N_j}) is the ( j) evaluation category, ( {C_i}) is the ( i) evaluation index, and ({V_{ij}} = left( {{a_{ij}},{b_{ij}}} right)left( {i = 1,2, cdots ,n;j = 1,2, cdots ,m} right)) is the range of quantity values ( {N_j}) for the index ( {C_i}), which is the classical domain of the data range taken by each category for the corresponding evaluation index.$$ {R_P} = left( {P,{C_i},{V_{iP}}} right) = left[ {begin{array}{*{20}{c}} P&{C_1}&{{V_{1P}}} \ {}&{C_2}&{{V_2}_P} \ {}& vdots & vdots \ {}&{C_n}&{{V_{nP}}} end{array}} right] = left[ {begin{array}{*{20}{c}} P&{C_1}&{left( {{a_{1P}},{b_{1P}}} right)} \ {}&{C_2}&{left( {{a_{2P}},{b_{2P}}} right)} \ {}& vdots & vdots \ {}&{C_n}&{({a_{nP}},{b_{nP}})} end{array}} right] $$
    (10)
    where ( P) is the whole of the category, ( {V_{iP}}) is the range of quantity values taken of ( P) for ( {C_i}), and ( {R_P}) is the ( P) joint domain.Determination of the matter element to be evaluatedFor ( q) to be evaluated and using the matter element to express the detected data or analysis results, the matter element ( {R_q}) to be evaluated can be expressed as$$ {R_q} = left( {q,{C_i},{v_i}} right) = left[ {begin{array}{*{20}{c}} q&{C_1}&{v_1} \ {}&{C_2}&{v_2} \ {}& vdots & vdots \ {}&{C_n}&{v_n} end{array}} right] $$
    (11)
    where ( q) is some thing and ( {v_i}) is the quantity value ( q) for ( {C_i}), which are the specific data obtained by the monitoring of the things that are to be evaluated.Determination and calculation of the degree of relationDetermination of the degree of relation for the thing to be evaluated in each category is expressed as follows:$$ {K_j}left( {v_i} right) = = left[ {begin{array}{*{20}{c}} {frac{{rho left( {{v_i},{V_{ij}}} right)}}{{rho left( {{v_i},{V_{iP}}} right) – rho left( {{v_i},{V_{ij}}} right)}}begin{array}{*{20}{c}} {}&{} end{array}rho left( {{v_i},{V_{iP}}} right) – rho left( {{v_i},{V_{ij}}} right) ne 0} \ {begin{array}{*{20}{c}} {}&{} end{array} – rho left( {{v_i},{V_{ij}}} right) – 1begin{array}{*{20}{c}} {}&{}&{} end{array}rho left( {{v_i},{V_{iP}}} right) – rho left( {{v_i},{V_{ij}}} right) = 0} end{array}} right] $$where ( rho left( {{v_i},{V_{ij}}} right) = rho left( {{v_i},left( {{a_{ij}},{b_{ij}}} right)} right) = left| {{v_i} – frac{{{a_{ij}} + {b_{ij}}}}{2}} right|-frac{{{b_{ij}} – {a_{ij}}}}{2}).The calculation of the thing ( q) to be evaluated for the degree of relation ( j) is expressed as$$ {K_j}left( q right) = sumlimits_{i = 1}^n {{a_i}{K_j}left( {v_i} right)} $$Determination of the levelDetermination of the level is expressed as follows:If ( {K_{j0}} = max left{ {K{}_jleft( q right)} right},j in left( {1,2, cdots ,m} right)), ( q) belongs to level ( {j_0}).In the extension set, the concept of a relational function is established. Any element in ( U) can be quantitatively described by the relational function value, which can belong to the positive, negative, or zero domains (i.e., belongs to the elements in the same domain). It is also possible to separate different levels from the size of the relational function valu27,28,29,30. More

  • in

    Distribution of trace elements in benthic infralittoral organisms from the western Antarctic Peninsula reveals no latitudinal gradient of pollution

    1.IUPAC. Compendium of Chemical Terminology, 2nd ed. (the ‘Gold Book’). Compiled by McNaught, A. D. & Wilkinson, A. (Blackwell Scientific Publications, 1997).2.Bolan, N. S., Adriano, D. C. & Naidu, R. Role of phosphorus in (im)mobilization and bioavailability of heavy metals in the soil-plant system. In Reviews of Environmental Contamination and Toxicology Vol. 177 (ed. Ware, G. W.) 1–44 (Springer, 2003).Chapter 

    Google Scholar 
    3.Marcovecchio, J., Botté, S., Domini, C. & Freije, R. Heavy metals, major metals, trace elements. In Handbook of Water Analysis (eds Nollet, L. M. L. & De Gelder, L. S. P.) 379–428 (CRC Press, 2013).
    Google Scholar 
    4.Wedepohl, H. K. The composition of the continental crust. Geochim. Cosmochim. Acta 59, 1217–1232 (1995).ADS 
    CAS 
    Article 

    Google Scholar 
    5.Santos, I. R., Silva-Filho, E. V., Schaefer, C. E. G. R., Albuquerque-Filho, M. R. & Campos, L. S. Heavy metal contamination in coastal sediments and soils near the Brazilian Antarctic Station, King George Island. Mar. Pollut. Bull. 50, 185–194 (2005).CAS 
    Article 

    Google Scholar 
    6.Khan, S., Cao, Q., Zheng, Y. M., Huang, Y. Z. & Zhu, Y. G. Health risks of heavy metals in contaminated soils and food crops irrigated with wastewater in Beijing, China. Environ. Pollut. 152, 686–692 (2008).CAS 
    Article 

    Google Scholar 
    7.Kabata-Pendias, A. Trace elements in soils and plants, 4th ed. (CRC Press, 2010).8.Waller, C. L. et al. Microplastics in the Antarctic marine system: An emerging area of research. Sci. Total Environ. 598, 220–227 (2017).ADS 
    CAS 
    Article 

    Google Scholar 
    9.Bargagli, R. Environmental contamination in Antarctic ecosystems. Sci. Total Environ. 400, 212–226 (2008).ADS 
    CAS 
    Article 

    Google Scholar 
    10.Lenihan, H. S., Oliver, J. S., Oakden, J. M. & Stephenson, M. D. Intense and localized benthic marine pollution around McMurdo Station, Antarctica. Mar. Pollut. Bull. 21, 422–430 (1990).CAS 
    Article 

    Google Scholar 
    11.Santos, I. R. et al. Baseline mercury and zinc concentrations in terrestrial and coastal organisms of Admiralty Bay, Antarctica. Environ. Pollut. 140, 304–311 (2006).Article 
    CAS 

    Google Scholar 
    12.Tin, T. et al. Impacts of local human activities on the Antarctic environment. Antarct. Sci. 21, 3–33 (2009).ADS 
    Article 

    Google Scholar 
    13.Corsolini, S. Industrial contaminants in Antarctic biota. J. Chromatogr. A 1216, 598–612 (2009).CAS 
    Article 

    Google Scholar 
    14.Bargagli, R., Agnorelli, C., Borghini, F. & Monaci, F. Enhanced deposition and bioaccumulation of mercury in antarctic terrestrial ecosystems facing a coastal polynya. Environ. Sci. Technol. 39, 8150–8155 (2005).ADS 
    CAS 
    Article 

    Google Scholar 
    15.Planchon, F. A. M. et al. Changes in heavy metals in Antarctic snow from Coats Land since the mid-19th to the late-20th century. Earth Planet. Sci. Lett. 200, 207–222 (2002).ADS 
    CAS 
    Article 

    Google Scholar 
    16.Szopińska, M., Namieśnik, J. & Polkowska, Ż How important is research on pollution levels in Antarctica? Historical approach, difficulties and current trends. In Reviews of Environmental Contamination and Toxicology Vol. 239 (ed. de Voogt, P.) 79–156 (Springer, 2017).
    Google Scholar 
    17.Bengtson Nash, S. et al. Contaminant profiles of air and soil around Casey station, Antarctica; discerning local and distant contaminant sources. In 21st Society for Environmental Toxicology and Chemistry (SETAC) Europe Annual Meeting Proceedings (2011).18.Boutron, C. F. & Patterson, C. C. Relative levels of natural and anthropogenic lead in recent Antarctic snow. J. Geophys. Res. 92, 8454–8464 (1987).ADS 
    CAS 
    Article 

    Google Scholar 
    19.Dick, A. L. Concentrations and sources of metals in the Antarctic Peninsula aerosol. Geochim. Cosmochim. Acta 55, 1827–1836 (1991).ADS 
    CAS 
    Article 

    Google Scholar 
    20.de Moreno, J. E. A., Gerpe, M. S., Moreno, V. J. & Vodopivez, C. Heavy metals in Antarctic organisms. Polar Biol. 17, 131–140 (1997).Article 

    Google Scholar 
    21.Kennicutt, I. M. C. et al. Human contamination of the marine environment-arthur harbor and mcmurdo sound, Antarctica. Environ. Sci. Technol. 29, 1279–1287 (1995).ADS 
    CAS 
    Article 

    Google Scholar 
    22.Hughes, K. A. & Ashton, G. V. Breaking the ice: The introduction of biofouling organisms to Antarctica on vessel hulls. Aquat. Conserv. Mar. Freshw. Ecosyst. 27, 158–164 (2017).Article 

    Google Scholar 
    23.Aston, S. R. & Thornton, I. Regional geochemical data in relation to seasonal variations in water quality. Sci. Total Environ. 7, 247–260 (1977).ADS 
    CAS 
    Article 

    Google Scholar 
    24.Norwood, W. P., Borgmann, U. & Dixon, D. G. Saturation models of arsenic, cobalt, chromium and manganese bioaccumulation by Hyalella azteca. Environ. Pollut. 143, 519–528 (2006).CAS 
    Article 

    Google Scholar 
    25.Jerez, S. et al. Concentration of trace elements in feathers of three Antarctic penguins: Geographical and interspecific differences. Environ. Pollut. 159, 2412–2419 (2011).CAS 
    Article 

    Google Scholar 
    26.Negri, A., Burns, K., Boyle, S., Brinkman, D. & Webster, N. Contamination in sediments, bivalves and sponges of McMurdo Sound, Antarctica. Environ. Pollut. 143, 456–467 (2006).CAS 
    Article 

    Google Scholar 
    27.Trevizani, T. H. et al. Bioaccumulation of heavy metals in marine organisms and sediments from Admiralty Bay, King George Island, Antarctica. Mar. Pollut. Bull. 106, 366–371 (2016).CAS 
    Article 

    Google Scholar 
    28.Trevizani, T. H., Petti, M. A. V., Ribeiro, A. P., Corbisier, T. N. & Figueira, R. C. L. Heavy metal concentrations in the benthic trophic web of Martel Inlet, Admiralty Bay (King George Island, Antarctica). Mar. Pollut. Bull. 130, 198–205 (2018).CAS 
    Article 

    Google Scholar 
    29.Cipro, C. V. Z., Montone, R. C. & Bustamante, P. Mercury in the ecosystem of Admiralty Bay, King George Island, Antarctica: Occurrence and trophic distribution. Mar. Pollut. Bull. 114, 564–570 (2017).CAS 
    Article 

    Google Scholar 
    30.de Oliveira, M. F. et al. Evidence of metabolic microevolution of the limpet Nacella concinna to naturally high heavy metal levels in Antarctica. Ecotoxicol. Environ. Saf. 135, 1–9 (2017).Article 
    CAS 

    Google Scholar 
    31.Torres, M. A. et al. Biochemical biomarkers in algae and marine pollution: A review. Ecotoxicol. Environ. Saf. 71, 1–15 (2008).CAS 
    Article 

    Google Scholar 
    32.Neff, J. M. Bioaccumulation in Marine Organisms. Effect of Contaminants from Oil Well Produced Water. Organic Geochemistry (Elsevier, 2002).
    Google Scholar 
    33.Wong, P. T. & Trevors, J. T. Chromium toxicity to algae and bacteria. In Chromium in the Natural and Human Environments (eds Nriagu, J. O. & Nieboer, E.) 305–315 (Wiley, 1988).
    Google Scholar 
    34.Community, E. Directive 2000/60/EC of the European Parliament and of the Council of 23 October 2000 establishing a framework for Community action in the field of water policy. Off. J. Eur. Parliam. L327, 1–82 (2000).
    Google Scholar 
    35.Driscoll, C. T., Mason, R. P., Chan, H. M., Jacob, D. J. & Pirrone, N. Mercury as a global pollutant: Sources, pathways, and effects. Environ. Sci. Technol. 47, 4967–4983 (2013).ADS 
    CAS 
    PubMed Central 
    Article 
    PubMed 

    Google Scholar 
    36.Pertierra, L. R. et al. Ecosystem services in Antarctica: Global assessment of the current state, future challenges and managing opportunities. Ecosyst. Serv. 49, 101299 (2021).Article 

    Google Scholar 
    37.Pringle, B. H., Hissong, D. E., Katz, E. L. & Mulawka, S. T. Trace metal accumulation by estuarine mollusks. J. Sanit. Eng. Div. Proc. Amer. Soc. Civ. Eng. 94, 455–475 (1968).CAS 
    Article 

    Google Scholar 
    38.Amiard, J. C., Amiard-Triquet, C., Berthet, B. & Metayer, C. Comparative study of the patterns of bioaccumulation of essential (Cu, Zn) and non-essential (Cd, Pb) trace metals in various estuarine and coastal organisms. J. Exp. Mar. Bio. Ecol. 106, 73–89 (1987).CAS 
    Article 

    Google Scholar 
    39.Borgmann, U., Norwood, W. P. & Clarke, C. Accumulation, regulation and toxicity of copper, zinc, lead and mercury in Hyalella azteca. Hydrobiologia 259, 79–89 (1993).CAS 
    Article 

    Google Scholar 
    40.Windom, H. & Kendall, D. R. Accumulation and biotransformation of mercury in coastal and marine biota. In The Biogeochemistry of Mercury in the Environment (ed. Nriagu, J. O.) 303–323 (Elsevier/North-Holland Biomedical Press, 1979).
    Google Scholar 
    41.Turner, S. J. et al. Are soft-sediment communities stable? An example from a windy harbour. Mar. Ecol. Prog. Ser. 120, 219–230 (1995).ADS 
    Article 

    Google Scholar 
    42.Caccia, V. G., Millero, F. J. & Palanques, A. The distribution of trace metals in Florida Bay sediments. Mar. Pollut. Bull. 46, 1420–1433 (2003).CAS 
    Article 

    Google Scholar 
    43.Gibbs, R. J. Transport phases of transition metals in the Amazon and Yukon Rivers. Bull. Geol. Soc. Am. 88, 829–843 (1977).CAS 
    Article 

    Google Scholar 
    44.Jain, C. K. & Sharma, M. K. Distribution of trace metals in the Hindon River system, India. J. Hydrol. 253, 81–90 (2001).ADS 
    CAS 
    Article 

    Google Scholar 
    45.Filgueiras, A. V., Lavilla, I. & Bendicho, C. Chemical sequential extraction for metal partitioning in environmental solid samples. J. Environ. Monit. 4, 823–857 (2002).CAS 
    Article 

    Google Scholar 
    46.Salomons, W. & Förstner, U. Metals in the Hydrocycle (Springer, 1984).Book 

    Google Scholar 
    47.Niimi, A. J. & Kissoon, G. P. Evaluation of the critical body burden concept based on inorganic and organic mercury toxicity to rainbow trout (Oncorhynchus mykiss). Arch. Environ. Contam. Toxicol. 26, 169–178 (1994).CAS 
    Article 

    Google Scholar 
    48.Landrum, P. F., Lydy, M. J. & Lee, H. Toxicokinetics in aquatic systems: Model comparisons and use in hazard assessment. Environ. Toxicol. Chem. 11, 1709–1725 (1992).CAS 
    Article 

    Google Scholar 
    49.Wiener, J. G. et al. Monitoring and evaluating trends in methylmercury accumulation in aquatic biota. In Ecosystem Responses to Mercury Contamination: Indicators of Change (eds Harris, R. et al.) 87–122 (CRC Press & SETAC Press, 2007).Chapter 

    Google Scholar 
    50.Dunton, K. H. δ15N and δ13C measurements of Antarctic Peninsula fauna: Trophic relationships and assimilation of benthic seaweeds. Am. Zool. 41, 99–112 (2001).
    Google Scholar 
    51.Corbisier, T. N., Petti, M. A. V., Skowronski, R. S. P. & Brito, T. A. S. Trophic relationships in the nearshore zone of Martel Inlet (King George Island, Antarctica): δ13C stable-isotope analysis. Polar Biol. 27, 75–82 (2004).Article 

    Google Scholar 
    52.Norkko, A. et al. Trophic structure of coastal Antarctic food webs associated with changes in sea ice and food supply. Ecology 88, 2810–2820 (2007).CAS 
    Article 

    Google Scholar 
    53.Michel, L. N. et al. Increased sea ice cover alters food web structure in East Antarctica. Sci. Rep. 9, 1–11 (2019).
    Google Scholar 
    54.Zenteno, L. et al. Unraveling the multiple bottom-up supplies of an Antarctic nearshore benthic community. Prog. Oceanogr. 174, 55–63 (2019).ADS 
    Article 

    Google Scholar 
    55.Cardona, L., Lloret-Lloret, E., Moles, J. & Avila, C. Latitudinal changes in the trophic structure of benthic coastal food webs in the Antarctic Peninsula. Mar. Environ. Res. 167, 105290 (2021).CAS 
    Article 

    Google Scholar 
    56.COMNAP. Antarctic Station Catalogue (COMNAP Secretariat, 2017).57.Wiencke, C., Amsler, C. & Clayton, M. Macroalgae. In Biogeographic Atlas of the Southern Ocean (eds De Broyer, C. et al.) 66–73 (Scientific Committee on Antarctic Research, 2014).
    Google Scholar 
    58.Danis, B., Griffiths, H. J. & Jangoux, M. Asteroidea. In Biogeographic Atlas of the Southern Ocean (eds De Broyer, C. et al.) 200–207 (Scientific Committee on Antarctic Research, 2014).
    Google Scholar 
    59.Schiaparelli, S. & Linse, K. Gastropoda. In Biogeographic Atlas of the Southern Ocean (eds De Broyer, C. et al.) 122–125 (Scientific Committee on Antarctic Research, 2014).
    Google Scholar 
    60.Borrell, A., Tornero, V., Bhattacharjee, D. & Aguilar, A. Trace element accumulation and trophic relationships in aquatic organisms of the Sundarbans mangrove ecosystem (Bangladesh). Sci. Total Environ. 545–546, 414–423 (2016).ADS 
    Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    61.Torres, J., Eira, C., Miquel, J. & Feliu, C. Heavy metal accumulation by intestinal helminths of vertebrates. Recent Adv. Pharm. Sci. II(661), 169–181 (2012).
    Google Scholar 
    62.Vighi, M., Borrell, A. & Aguilar, A. Bone as a surrogate tissue to monitor metals in baleen whales. Chemosphere 171, 81–88 (2017).ADS 
    CAS 
    Article 

    Google Scholar 
    63.Borrell, A., Aguilar, A., Tornero, V. & Drago, M. Concentrations of mercury in tissues of striped dolphins suggest decline of pollution in Mediterranean open waters. Chemosphere 107, 319–323 (2014).ADS 
    CAS 
    Article 

    Google Scholar 
    64.Borrell, A., Clusa, M., Aguilar, A. & Drago, M. Use of epidermis for the monitoring of tissular trace elements in Mediterranean striped dolphins (Stenella coeruleoalba). Chemosphere 122, 288–294 (2015).ADS 
    CAS 
    Article 

    Google Scholar 
    65.Maceda-Veiga, A., Monroy, M., Navarro, E., Viscor, G. & de Sostoa, A. Metal concentrations and pathological responses of wild native fish exposed to sewage discharge in a Mediterranean river. Sci. Total Environ. 449, 9–19 (2013).ADS 
    CAS 
    Article 

    Google Scholar 
    66.Suda, C. N. K. et al. The biology and ecology of the Antarctic limpet Nacella concinna. Polar Biol. 38, 1949–1969 (2015).Article 

    Google Scholar 
    67.Škrbić, B., Crossed, D. & Signurišić-Mladenović, N. Distribution of heavy elements in urban and rural surface soils: The Novi Sad city and the surrounding settlements. Serbia. Environ. Monit. Assess. 185, 457–471 (2013).Article 
    CAS 

    Google Scholar 
    68.Škrbić, B. D., Buljovčić, M., Jovanović, G. & Antić, I. Seasonal, spatial variations and risk assessment of heavy elements in street dust from Novi Sad, Serbia. Chemosphere 205, 452–462 (2018).ADS 
    Article 
    CAS 

    Google Scholar 
    69.Škrbić, B., Durišić-Mladenović, N. & Cvejanov, J. Principal component analysis of trace elements in Serbian wheat. J. Agric. Food Chem. 53, 2171–2175 (2005).Article 
    CAS 

    Google Scholar 
    70.Wilde, E. W. & Benemann, J. R. Bioremoval of heavy metals by the use of microalgae. Biotechnol. Adv. 11, 781–812 (1993).CAS 
    Article 

    Google Scholar 
    71.Farías, S., Arisnabarreta, S. P., Vodopivez, C. & Smichowski, P. Levels of essential and potentially toxic trace metals in Antarctic macro algae. Spectrochim. Acta B 57, 2133–2140 (2002).ADS 
    Article 

    Google Scholar 
    72.Black, W. A. P. & Mitchell, R. L. Trace elements in the common algae and in sea water. J. Mar. Biol. Assoc. UK 30, 1–10 (1952).Article 

    Google Scholar 
    73.Lignell, A., Roomans, G. M. & Pedersen, M. Localization of absorbed cadmium in Fucus vesiculosus L. by X-ray microanalysis. Z. Pflanzenphysiol. 105, 103–109 (1982).CAS 
    Article 

    Google Scholar 
    74.Ragan, M. A., Smidsrød, O. & Larsen, B. Chelation of divalent metal ions by brown algal polyphenols. Mar. Chem. 7, 265–271 (1979).CAS 
    Article 

    Google Scholar 
    75.Talarico, L. Fine structure and X-ray microanalysis of a red macrophyte cultured under cadmium stress. Environ. Pollut. 120, 813–821 (2002).CAS 
    Article 

    Google Scholar 
    76.Vasconcelos, M. T. S. D. & Leal, M. F. C. Seasonal variability in the kinetics of Cu, Pb, Cd and Hg accumulation by macroalgae. Mar. Chem. 74, 65–85 (2001).CAS 
    Article 

    Google Scholar 
    77.Pellegrini, L., Delivopoulos, S. G. & Pellegrini, M. Arsenic-induced ultrastructural changes in the vegetative cells of Cystoseira barbata forma repens Zinova et Kalugina (Fucophyceae, Fucales). Bot. Mar. 33, 229–234 (1990).CAS 
    Article 

    Google Scholar 
    78.Deheyn, D. D., Gendreau, P., Baldwin, R. J. & Latz, M. I. Evidence for enhanced bioavailability of trace elements in the marine ecosystem of Deception Island, a volcano in Antarctica. Mar. Environ. Res. 60, 1–33 (2005).CAS 
    Article 

    Google Scholar 
    79.Exley, C. Silicon in life: A bioinorganic solution to bioorganic essentiality. J. Inorg. Biochem. 69, 139–144 (1998).CAS 
    Article 

    Google Scholar 
    80.Costa, R. R. et al. Dynamics of an intense diatom bloom in the Northern Antarctic Peninsula, February 2016. Limnol. Oceanogr. 65, 2056–2075 (2020).ADS 
    CAS 
    Article 

    Google Scholar 
    81.Mendes, C. R. B. et al. Dynamics of phytoplankton communities during late summer around the tip of the Antarctic Peninsula. Deep. Res. I 65, 1–14 (2012).CAS 
    Article 

    Google Scholar 
    82.Ducklow, H. W. et al. Marine pelagic ecosystems: The West Antarctic Peninsula. Philos. Trans. R. Soc. B. 362, 67–94 (2007).Article 

    Google Scholar 
    83.Prezelin, B. B., Hofmann, E. E., Mengelt, C. & Klinck, J. M. The linkage between Upper Circumpolar Deep Water (UCDW) and phytoplankton assemblages on the west Antarctic Peninsula continental shelf. J. Mar. Res. 58, 165–202 (2000).Article 

    Google Scholar 
    84.Bargagli, R., Monaci, F. & Cateni, D. Marine coastal food web. Mar. Ecol. Prog. Ser. 169, 65–76 (1998).ADS 
    CAS 
    Article 

    Google Scholar 
    85.Collier, R. & Edmond, J. The trace element geochemistry of marine biogenic particulate matter. Prog. Oceanogr. 13, 113–199 (1984).ADS 
    Article 

    Google Scholar 
    86.Rubio, C. et al. Metals in edible seaweed. Chemosphere 173, 572–579 (2017).ADS 
    CAS 
    Article 

    Google Scholar 
    87.Desideri, D. et al. Essential and toxic elements in seaweeds for human consumption. J. Toxicol. Environ. Health A 79, 112–122 (2016).CAS 
    Article 

    Google Scholar 
    88.Runcie, J. W. & Riddle, M. J. Metal concentrations in macroalgae from East Antarctica. Mar. Pollut. Bull. 49, 1114–1119 (2004).CAS 
    Article 

    Google Scholar 
    89.Fowler, S. W., Villeneuve, J. P., Wyse, E., Jupp, B. & de Mora, S. Temporal survey of petroleum hydrocarbons, organochlorinated compounds and heavy metals in benthic marine organisms from Dhofar, southern Oman. Mar. Pollut. Bull. 54, 357–367 (2007).CAS 
    Article 

    Google Scholar 
    90.Curtosi, A., Pelletier, E., Vodopivez, C., St Louis, R. & MacCormack, W. P. Presence and distribution of persistent toxic substances in sediments and marine organisms of Potter Cove, Antarctica. Arch. Environ. Contam. Toxicol. 59, 582–592 (2010).CAS 
    Article 

    Google Scholar 
    91.Ahn, I. Y., Kim, K. W. & Choi, H. J. A baseline study on metal concentrations in the Antarctic limpet Nacella concinna (Gastropoda: Patellidae) on King George Island: Variations with sex and body parts. Mar. Pollut. Bull. 44, 424–431 (2002).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    92.Dayton, P. K., Robilliard, G. A., Paine, R. T. & Dayton, L. B. Biological accommodation in the benthic community at McMurdo Sound, Antarctica. Ecol. Monogr. 44, 105–128 (1974).Article 

    Google Scholar 
    93.Pearse, J. S. Reproductive periodicities in several contrasting populations of Odontaster validus Koehler, a common Antarctic
    asteroid. Antarct. Res. Ser. 5, 39–85 (1965).94.Peckham, V. Year-round SCUBA diving in the Antarctic. Polar Rec. 12, 143–146 (1964).Article 

    Google Scholar 
    95.Smale, D. A., Barnes, D. K. A., Fraser, K. P. P., Mann, P. J. & Brown, M. P. Scavenging in Antarctica: Intense variation between sites and seasons in shallow benthic necrophagy. J. Exp. Mar. Bio. Ecol. 349, 405–417 (2007).Article 

    Google Scholar 
    96.Mcclintock, J. B. Trophic biology of antarctic shallow-water echinoderms. Mar. Ecol. Prog. Ser. 111, 191–202 (1994).ADS 
    Article 

    Google Scholar 
    97.Grotti, M. et al. Natural variability and distribution of trace elements in marine organisms from Antarctic coastal environments. Antarct. Sci. 20, 39–51 (2008).ADS 
    Article 

    Google Scholar 
    98.Papadopoulou, C., Kanias, G. D. & Moraitopoulou-Kassimati, E. Stable elements of radioecological importance in certain echinoderm species. Mar. Pollut. Bull. 7, 143–144 (1976).CAS 
    Article 

    Google Scholar 
    99.Di Giglio, S. et al. Effects of ocean acidification on acid-base physiology, skeleton properties, and metal contamination in two echinoderms from vent sites in Deception Island, Antarctica. Sci. Total Environ. 765, 142669 (2020).Article 
    CAS 

    Google Scholar 
    100.Danis, B. et al. Contaminant levels in sediments and asteroids (Asterias rubens L., Echinodermata) from the Belgian coast and Scheldt estuary: Polychlorinated biphenyls and heavy metals. Sci. Total Environ. 333, 149–165 (2004).ADS 
    CAS 
    Article 

    Google Scholar 
    101.Riva, S. D., Abelmoschi, M. L., Magi, E. & Soggia, F. The utilization of the Antarctic environmental specimen bank (BCAA) in monitoring Cd and Hg in an Antarctic coastal area in Terra Nova Bay (Ross Sea: Northern Victoria Land). Chemosphere 56, 59–69 (2004).ADS 
    Article 
    CAS 

    Google Scholar 
    102.Cabrita, M. T. et al. Evaluating trace element bioavailability and potential transfer into marine food chains using immobilised diatom model species Phaeodactylum tricornutum, on King George Island, Antartica. Mar. Pollut. Bull. 121, 192–200 (2017).CAS 
    Article 

    Google Scholar 
    103.Truzzi, C. et al. Separation of micro-phytoplankton from inorganic particulate in Antarctic seawater (Ross Sea) for the determination of Cd, Pb and Cu: Optimization of the analytical methodology. Anal. Methods 7, 5490–5496 (2015).Article 

    Google Scholar 
    104.Bargagli, R. Trace metals in Antarctic organisms and the development of circumpolar biomonitoring networks. in Reviews of Environmetal Contamination and Toxicology (ed. Ware, G. W.) vol. 171, 53–110 (2001).105.Focardi, S., Bargagli, R. & Corsolini, S. Isomer-specific analysis and toxic potential evaluation of polychlorinated biphenyls in Antarctic fish, seabirds and Weddell seals from Terra Nova Bay (Ross Sea). Antarct. Sci. 7, 31–35 (1995).ADS 
    Article 

    Google Scholar 
    106.Demina, L. L. & Nemirovskaya, I. A. Spatial distribution of microelements in the seston of the White Sea. Oceanology 47, 360–372 (2007).ADS 
    Article 

    Google Scholar 
    107.Wiencke, C. & Amsler, C. D. Seaweeds and their communities in polar regions. In Seaweed Biology (eds Wiencke, C. & Bischof, K.) 265–291 (Springer, 2012).Chapter 

    Google Scholar 
    108.Fairhead, V. A., Amsler, C. D., Mcclintock, J. B. & Baker, B. J. Within-thallus variation in chemical and physical defences in two species of ecologically dominant brown macroalgae from the Antarctic Peninsula. Oceanology 322, 1–12 (2005).CAS 

    Google Scholar 
    109.Amsler, C. D. Algal chemical ecology: Algal Chemical Ecology (Springer, 2008).Book 

    Google Scholar 
    110.Amsler, C. D., Mcclintock, J. B. & Baker, B. J. Chemical mediation of mutualistic interactions between macroalgae and mesograzers structure unique coastal communities along the western Antarctic Peninsula. J. Phycol. 50, 1–10 (2014).Article 

    Google Scholar 
    111.Aumack, C. F., Amsler, C. D., McClintock, J. B. & Baker, B. J. Chemically mediated resistance to mesoherbivory in finely branched macroalgae along the western Antarctic Peninsula. Eur. J. Phycol. 45, 19–26 (2010).Article 

    Google Scholar 
    112.Núñez-Pons, L., Rodríguez-Arias, M., Gómez-Garreta, A., Ribera-Siguán, A. & Avila, C. Feeding deterrency in Antarctic marine organisms: Bioassays with the omnivore amphipod Cheirimedon femoratus. Eur. J. Phycol. 462, 163–174 (2012).
    Google Scholar 
    113.Ahn, I. Y., Chung, K. H. & Choi, H. J. Influence of glacial runoff on baseline metal accumulation in the Antarctic limpet Nacella concinna from King George Island. Mar. Pollut. Bull. 49, 119–127 (2004).CAS 
    Article 

    Google Scholar 
    114.Burdon-Jones, C., Denton, G. R. W., Jones, G. B. & McPhie, K. A. Regional and seasonal variations of trace metals in tropical phaeophyceae from North Queensland. Mar. Environ. Res. 7, 13–30 (1982).CAS 
    Article 

    Google Scholar 
    115.Augier, H., Gilles, G., Leal Nascimento, M. & Ramonda, G. Évolution de la contamination de la flore et de la faune marines benthiques de la Baie de Port-Cros de 1976 à 1981. Trav. Sci. Parc Natl. Port-Cros 10, 37–50 (1984).
    Google Scholar 
    116.Chakraborty, S., Bhattacharya, T., Singh, G. & Maity, J. P. Benthic macroalgae as biological indicators of heavy metal pollution in the marine environments: A biomonitoring approach for pollution assessment. Ecotoxicol. Environ. Saf. 100, 61–68 (2014).CAS 
    Article 

    Google Scholar 
    117.Pastor, A. et al. Levels of heavy metals in some marine organisms from the western Mediterranean area (Spain). Mar. Pollut. Bull. 28, 50–53 (1994).CAS 
    Article 

    Google Scholar  More

  • in

    Global variation in the fraction of leaf nitrogen allocated to photosynthesis

    In this study, we produced a global fLNR and ({V}_{{c}_{{max }}}^{25}) map using an RF model trained primarily by remote sensing and in situ observations and examined seven ({V}_{{c}_{{max }}}^{25}) models based on 5 competing hypotheses with regard to their assumptions on fLNR. Our results suggested that the global average fLNR was 18.2 ± 6.2%, and the global distribution of fLNR was dominated by the interaction between fLNR and leaf traits (i.e., LMA and LPC), followed by regional influences from climate (i.e., VPD and PAR) and soil characteristics (i.e., soil pH and sand percentage). We used RF fLNR distribution and its relationships with environmental covariates to evaluate five empirical and two optimal ({V}_{{c}_{{max }}}^{25}) models, and found that the models showed different degrees of inefficacy in reproducing RF fLNR. Here, we discuss the mechanisms underlying the detected fLNR responses to leaf traits, climate, and soil characteristics and propose future directions to improve the simulation of fLNR and ({V}_{{c}_{{max }}}^{25}) in models.Negative correlation between fLNR and LMAOur finding that fLNR is negatively related to LMA agrees with a previous meta-analysis that found fLNR decreases by 0.54 ± 0.08% with a 1 g/m2 increase in LMA based on a univariate regression10, though another study reported that the negative relationship between fLNR and LMA was non-significant using a smaller dataset11. Using the global dataset, we found a relatively small sensitivity of fLNR to LMA (−0.19 ± 0.001% per 1 g/m2) when accounting for climate and soil (Fig. 3b).Higher LMA is the result of plants allocating more biomass and nitrogen to building cell walls, which may cause a reduction in CO2 diffusion into the mesophyll as well as relative nitrogen allocated to RuBisCO38. Leaves with greater LMA are tougher and usually have a longer leaf lifespan11,36. Therefore, the negative correlation between fLNR and LMA highlights the trade-off between photosynthesis and persistence along the leaf economic spectrum: on one end, leaves invest more nitrogen in RuBisCO to increase the photosynthetic capacity and enhance carbon uptake; on the other end leaves invest more nitrogen in structural biomass to improve leaf longevity and lengthen the carbon uptake period. The latter is especially true for evergreen species that have greater LMA and smaller fLNR than deciduous and herbaceous species10. The coordination of fLNR and LMA is also consistent with a recent analysis highlighting the role of LMA in determining the variation and predictability of LNC in ecosystem models39.In addition, we found that LPC increases fLNR in tropical evergreen forests and mixed forests, which tend to be more phosphorus limited40. Our result is consistent with previous studies reporting coupled leaf photosynthetic capacity (i.e., ({V}_{{c}_{{max }}}^{25}) or maximum photosynthetic capacity (Amax)) and LPC for tropical species41,42. This result indicates potential widespread adjustments of plants nitrogen use by phosphorus investment for photosynthesis and plant growth43 in tropical and mixed forests. In addition, we note that the productivity of some grasslands44,45 and boreal forests46,47 has also been reported to be limited by phosphorus availability, however, we did not detect a strong positive dependence of fLNR on LPC globally for these ecosystems in our study. The difference potentially suggests that the phosphorus limitation of grasslands and boreal forests is not as prevalent as that for tropical and mixed forests (though some mixed forests are in the boreal region).Climate and soil impacts on fLNRThe response of fLNR to climate is often implicitly included in ({V}_{{c}_{{max }}}^{25}) models. We found that fLNR was sensitive to annual VPD globally. Several studies have reported that plants in arid environments (i.e., high VPD) tend to have a higher Amax and LNC48,49 as plants enhance photosynthetic capacity to maintain a given assimilation rate with lower stomatal conductance and reduced water loss. Such a response to aridity has been described using the least-cost theory19,21. Our results show that other than Amax and LNC, fLNR also increases with VPD, consistent with a recent study reporting higher nutrient use efficiency for plants in semi-arid ecosystems of the African Sahel49. We note that an earlier study reporting differently that a dry site has a smaller ({V}_{{c}_{{max }}}^{25})/LNC ratio (i.e., smaller fLNR) than a wet site19, though it used annual precipitation, not VPD to define aridity.In addition, the positive relationship between PAR and fLNR for non-forests (Fig. 3c) provides a potential explanation of the light acclimation of photosynthesis, as several studies have found that leaf and ecosystem Amax can be enhanced by intermediate to long-term average PAR50,51,52. For non-forest ecosystems, our results suggest that photosynthetic light acclimation emerges as plants increase fLNR in response to increasing annual PAR. However, for forests (except EBF) the results suggest that photosynthetic light acclimation may emerge more due to the increase in LNC as we did not detect a positive response of fLNR to light (Fig. 3c).Soil characteristics have been reported to influence Amax and LNC37, but we found no studies that have examined the impact of soil characteristics on fLNR. Among the eight soil properties we examined, we found positive responses of fLNR to soil pH and soil sand percentage, followed by small influences of bulk density and silt for certain ecosystems (i.e., croplands, needle leaf forests). pH influences the ability of soil to hold on to nutrients, including Ca2+, K2+, and Mg2+, that are essential to plant growth. A higher pH means more available nutrient cations as acid soils replace nutrient cations with H+. Several studies have reported a positive effect of pH on Amax37, non-temperature standardized (V_{c_{max}})20, and LNC39. Soil sand percentage had a positive impact on fLNR, possibly because sandy soils tend to be less fertile53 and thus stimulate plants to use their nitrogen more efficiently for photosynthesis and growth. The global influence of soil on fLNR was generally smaller than leaf traits and climate, but our analysis indicated that on 11.9% of the vegetated surface, soil characteristics contributed more than 15% of the changes in fLNR (Fig. 3a).Notably, our study found that the soil nitrogen content has a limited impact on the spatial variation of fLNR (Fig. 3). The result implies that processes such as nitrogen deposition/addition are unlikely to affect plants fLNR. The soil nitrogen map we used was upscaled from ground observations of soil profiles in the World Soil Information Service (WoSIS) database. About 47.4–81.4% of the soil profiles in WoSIS are collected from the 1980s to 2020s54, when there were strong N deposition effects55. Therefore, we expect the N deposition effect has been implicitly included in our analysis. We acknowledge that some studies have suggested N deposition influenced leaf nitrogen content and photosynthesis56,57, however, the influence is limited to certain biomes, deposition load range, and time after the deposition. It is unclear whether these localized and time-dependent effects can influence the global variation of fLNR.Uncertainty in the derivation of fLNR and ({V}_{{c}_{{max}}})
    fLNR was derived based on Eq. (1) (see “Methods”) that mechanistically links ({V}_{{c}_{{max }}}^{25}), LNC, and fLNR, with the assumption that specific activity of RuBisCO (α25) and mass ratio of RuBisCO to nitrogen (fNR) are relatively constant values. The average uncertainty of RF fLNR was about 4.20 ± 2.20% (Supplementary Fig. 3). The uncertainty of fLNR was propagated from several sources including RF ({V}_{{c}_{{ma}x}}^{25}), α25, fNR, and LNC (Supplementary Fig. 3). Among them, the α25 ranges between 47.34 and 60 μmol CO2/g RuBisCO/s, and fNR ranges between 6.11 and 7.16 g RuBisCO/g N4. Our uncertainty test showed that the influence of α25 and fNR uncertainties on global fLNR were only around 1.13 ± 0.39% and 0.80 ± 0.27%, respectively (see “Methods”; Supplementary Fig. 3). Physiologically, α25 is a value that reflects the change in active sites of RuBisCO and the kinetic constant of the enzyme RuBisCO (k25). The number of active sites of RuBisCO is often regarded as a fixed value (set at 6 × 1023/mol RuBisCO) for vegetation on the land surface5, but there are reports showing that k25 varies with species9, leaf ages58, and temperature59. While these dependencies are elusive due to limited observations, previous studies have reported that k25 negatively correlates with LNC60 and LMA61. The negative relationship between k25 and LMA or LNC is potentially caused by the relatively lower drawdown of CO2 from intercellular spaces to the chloroplast as increased LMA increases mesophyll resistance. In that case, the negative dependence of k25 and α25 on LNC and LMA might account for part of the negative dependence of fLNR on LMA that we found (Fig. 3b), though the negative influence of LMA on α25 was weak and within the range of uncertainty, we quantified (Supplementary Fig. 3).Compared to α25 and fNR, the uncertainties in LNC and RF ({V}_{{c}_{{max }}}^{25}) incurred larger uncertainties in fLNR. We found that LNC alone caused changes of 3.35 ± 2.16% in fLNR and RF ({V}_{{c}_{{max }}}^{25}) caused 3.13 ± 1.50% (Supplementary Fig. 3). Our study is the first attempt to upscale in situ ({V}_{{c}_{{max }}}^{25}) to the globe using remote sensing, while similar studies have done that for other leaf traits33. The observations used for training RF were densely distributed in Europe and North America, while inner Asia, Southeast Asia, Africa, and high-latitude regions are much less constrained by observations (Supplementary Fig. 6a). In addition, we did not consider temperature acclimation when standardizing in situ (V_{c_{max}}) to ({V}_{{c}_{{max }}}^{25}) (Eq. (2)), in order to facilitate the comparison with models that only estimate ({V}_{{c}_{{max }}}^{25}). However, the uncertainty related to temperature scaling should be limited as acclimated and non-acclimated temperature scaling factors for (V_{c_{max}}) are similar under 30 °C62,63.The choice of an LNC map is another source of uncertainty in the derivation of fLNR. There are several global LNC maps available other than the EB1728 map we used, namely AMM1833 and CB2031. Each product has been validated in their respective studies (Supplementary Table 3). To examine the uncertainty incurred by the choice of LNC maps, we calculated fLNR using each of the three LNC maps. The three resulting fLNR maps show similar spatial patterns (Supplementary Fig. 10), with the spatial correlation coefficients (r) between them ranging from 0.57 to 0.71 (p  More

  • in

    Allopatric humpback whales of differing generations share call types between foraging and wintering grounds

    1.Seyfarth, R. M. & Cheney, D. L. Production, usage, and comprehension in animal vocalizations. Brain Lang. 115, 92–100 (2010).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    2.Forstmeier, W., Burger, C., Temnow, K. & Derégnaucourt, S. The genetic basis of zebra finch vocalizations. Evolution 63, 2114–2130 (2009).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    3.Kroodsma, D. E. & Konishi, M. A suboscine bird (eastern phoebe, Sayornis phoebe) develops normal song without auditory feedback. Anim. Behav. 42, 477–487 (1991).Article 

    Google Scholar 
    4.Crance, J. L., Bowles, A. E. & Garver, A. Evidence for vocal learning in juvenile male killer whales, Orcinus orca, from an adventitious cross-socializing experiment. J. Exp. Biol. 217, 1229–1237 (2014).PubMed 

    Google Scholar 
    5.Ralls, K., Fiorelli, P. & Gish, S. Vocalizations and vocal mimicry in captive harbor seals, Phoca vitulina. Can. J. Zool. 63, 1050–1056 (1985).Article 

    Google Scholar 
    6.Boughman, J. W. Vocal learning by greater spear-nosed bats. Proc. R. Soc. B Biol. Sci. 265, 227–233 (1998).CAS 
    Article 

    Google Scholar 
    7.Foote, A. D. et al. Killer whales are capable of vocal learning. Biol. Lett. 2, 509–512 (2006).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    8.Jones, G. & Ransome, R. D. Echolocation calls of bats are influenced by maternal effects and change over a lifetime. Proc. R. Soc. B Biol. Sci. 252, 125–128 (1993).ADS 
    CAS 
    Article 

    Google Scholar 
    9.Rendell, L. & Whitehead, H. Culture in whales and dolphins. Behav. Brain Sci. 24, 309–382 (2001).CAS 
    PubMed 
    Article 

    Google Scholar 
    10.Deecke, V. B., Ford, J. K. B. & Spong, P. Dialect change in resident killer whales: Implications for vocal learning and cultural transmission. Anim. Behav. 60, 629–638 (2000).CAS 
    PubMed 
    Article 

    Google Scholar 
    11.Filatova, O. A., Burdin, A. M. & Hoyt, E. Horizontal transmission of vocal traditions in killer whale (Orcinus orca) dialects. Biol. Bull. 37, 965–971 (2010).Article 

    Google Scholar 
    12.Garland, E. C. et al. Dynamic horizontal cultural transmission of humpback whale song at the ocean basin scale. Curr. Biol. 21, 687–691 (2011).CAS 
    PubMed 
    Article 

    Google Scholar 
    13.Proppe, D. S. et al. Black-capped chickadees Poecile atricapillus sing at higher pitches with elevated anthropogenic noise, but not with decreasing canopy cover. J. Avian Biol. 43, 325–332 (2012).Article 

    Google Scholar 
    14.Parks, S. E., Clark, C. W. & Tyack, P. L. Short- and long-term changes in right whale calling behavior: The potential effects of noise on acoustic communication. J. Acoust. Soc. Am. 122, 3725–3731 (2007).ADS 
    PubMed 
    Article 

    Google Scholar 
    15.Caldwell, M. C. & Caldwell, D. K. Individualized whistle contours in bottlenosed dolphins (Tursiops truncatus). Nature 207, 434–435 (1965).ADS 
    Article 

    Google Scholar 
    16.Waser, P. M. The evolution of male loud calls among mangabeys and baboons. In Primate communication (ed. Snowdon, C. T.) 117–143 (Cambridge University Press, 1982).
    Google Scholar 
    17.Payne, K. & Payne, R. Large scale changes over 19 years in songs of humpback whales in Bermuda. Z. Tierpsychol. 68, 89–114 (1985).Article 

    Google Scholar 
    18.Stimpert, A. K., Wiley, D. N., Au, W. W. L., Johnson, M. P. & Arsenault, R. ‘Megapclicks’: Acoustic click trains and buzzes produced during night-time foraging of humpback whales (Megaptera novaeangliae). Biol. Lett. 3, 467–470 (2007).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    19.Fournet, M. E. H., Gabriele, C. M., Sharpe, F., Straley, J. M. & Szabo, A. Feeding calls produced by solitary humpback whales. Mar. Mammal Sci. 1, 1–15 (2018).
    Google Scholar 
    20.Sloan, J. L., Wilson, D. R. & Hare, J. F. Functional morphology of Richardson’s ground squirrel, Spermophilus richardsonii, alarm calls: The meaning of chirps, whistles and chucks. Anim. Behav. 70, 937–944 (2005).Article 

    Google Scholar 
    21.Luther, D. & Baptista, L. Urban noise and the cultural evolution of bird songs. Proc. R. Soc. B 277, 469–473 (2010).PubMed 
    Article 

    Google Scholar 
    22.Weilgart, L. S. The impacts of anthropogenic ocean noise on cetaceans and implications for management. Can. J. Zool. 85, 1091–1116 (2007).Article 

    Google Scholar 
    23.Strager, H. Pod-specific call repertoires and compound calls of killer whales, Orcinus orca Linnaeus, 1758, in the waters of northern Norway. Can. J. Zool. 73, 1037–1047 (1995).Article 

    Google Scholar 
    24.Rehn, N., Filatova, O. A., Durban, J. W. & Foote, A. D. Cross-cultural and cross-ecotype production of a killer whale ‘excitement’ call suggests universality. Naturwissenschaften 98, 1–6 (2011).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    25.Fournet, M. E. H., Jacobsen, L., Gabriele, C. M., Mellinger, D. K. & Klinck, H. More of the same: Allopatric humpback whale populations share acoustic repertoire. PeerJ 6, e5365 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    26.Miksis-Olds, J. L., Harris, D. V. & Heaney, K. D. Comparison of estimated 20-Hz pulse fin whale source levels from the tropical Pacific and Eastern North Atlantic Oceans to other recorded populations. J. Acoust. Soc. Am. 146, 2373–2384 (2019).ADS 
    PubMed 
    Article 

    Google Scholar 
    27.Ford, J. K. B. Acoustic behaviour of resident killer whales (Orcinus orca) off Vancouver Island, British Columbia. Can. J. Zool. 67, 727–745 (1989).Article 

    Google Scholar 
    28.Ford, J. K. B. Vocal traditions among resident killer whales (Orcinus orca) in coastal waters of British Columbia. Can. J. Zool. 69, 1454–1483 (1991).Article 

    Google Scholar 
    29.Foote, A. D., Osborne, R. W. & Rus Hoelzel, A. Temporal and contextual patterns of killer whale (Orcinus orca) call type production. Ethology 114, 599–606 (2008).Article 

    Google Scholar 
    30.Terhune, J. Geographical variation of harp seal underwater vocalizations. Can. J. Zool. 72, 892–897 (1994).Article 

    Google Scholar 
    31.Serrano, A. & Terhune, J. M. Stability of the underwater vocal repertoire of harp seals (Pagophilus groenlandicus). Aquat. Mamm. 28, 1 (2002).
    Google Scholar 
    32.Risch, D. et al. Vocalizations of male bearded seals, Erignathus barbatus: Classification and geographical variation. Anim. Behav. 73, 747–762 (2007).Article 

    Google Scholar 
    33.Sayigh, L. S. et al. Individual recognition in wild bottlenose dolphins: A field test using playback experiments. Anim. Behav. 57, 41–50 (1998).Article 

    Google Scholar 
    34.Baker, C. S. et al. Migratory movement and population structure of humpback whales (Megaptera novaeangliae) in the central and eastern North Pacific. Mar. Ecol. Prog. Ser. 31, 105–119 (1986).ADS 
    Article 

    Google Scholar 
    35.Acevedo, J., Mora, C. & Aguayo-Lobo, A. Sex-related site fidelity of humpback whales (Megaptera novaeangliae) to the Fueguian Archipelago feeding area, Chile. Mar. Mammal Sci. 30, 433–444 (2014).Article 

    Google Scholar 
    36.Gabriele, C. M. et al. Natural history, population dynamics, and habitat use of humpback whales over 30 years on an Alaska feeding ground. Ecosphere 8, 1–10 (2017).Article 

    Google Scholar 
    37.Chittleborough, R. G. Dynamics of two populations of the humpback whale, Megaptera novaeangliae (Borowski). Aust. J. Mar. Freshwat. Res.16, 33–128 (1965).38.Baker, C. S. et al. Strong maternal fidelity and natal philopatry shape genetic structure in North Pacific humpback whales. Mar. Ecol. Prog. Ser. 494, 291–306 (2013).ADS 
    Article 

    Google Scholar 
    39.Valsecchi, E. et al. Microsatellite genetic distances between oceanic populations of the humpback whale (Megaptera novaeangliae). Mol. Biol. Evol. 14, 355–362 (1997).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    40.Jackson, J. A. et al. Global diversity and oceanic divergence of humpback whales (Megaptera novaeangliae). Proc. R. Soc. B Biol. Sci. 281, 20133222–20133222 (2014).Article 

    Google Scholar 
    41.Baker, C. S. et al. Abundant mitochondrial DNA variation and world-wide population structure in humpback whales. Proc. Natl. Acad. Sci. USA. 90, 8239–8243 (1993).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    42.Dawbin, W. H. The seasonal migratory cycle of humpback whales. In Whales, dolphins and porpoises (ed. Norris, K. S.) 145–171 (University of California Press, 1966).Chapter 

    Google Scholar 
    43.Baker, C. S. & Herman, L. M. Seasonal contrasts in the social behavior of the humpback whale. CETUS 5, 14–16 (1984).
    Google Scholar 
    44.D’Vincent, C. G., Nilson, R. M. & Hanna, R. E. Vocalization and coordinated feeding behavior of the humpback whale in southeastern Alaska. Sci. Rep. Whale Res. Inst. Tokyo 1, 41–47 (1985).
    Google Scholar 
    45.Baraff, L. S., Clapham, P. J., Mattila, D. & Bowman, R. S. Feeding behaviour of a humpback whale in low-latitudes. Mar. Mammal Sci. 7, 197–202 (1991).Article 

    Google Scholar 
    46.Silber, G. K. The relationship of social vocalizations to surface behavior and aggression in the Hawaiian humpback whale (Megaptera novaeangliae). Can. J. Zool. 64, 2075–2080 (1986).Article 

    Google Scholar 
    47.Tyack, P. & Whitehead, H. Male competition in large groups of wintering humpback whales. Behaviour 83, 132–154 (1982).Article 

    Google Scholar 
    48.Baker, C. S. & Herman, L. M. Aggressive behavior between humpback whales (Megaptera novaeangliae) wintering in Hawaiian waters. Can. J. Zool. 62, 1922–1937 (1984).Article 

    Google Scholar 
    49.Payne, R. S. & McVay, S. Songs of humpback whales. Science 173, 585–597 (1971).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    50.Cerchio, S., Jacobsen, J. K. & Norris, T. F. Temporal and geographical variation in songs of humpback whales, Megaptera novaeangliae: Synchronous change in Hawaiian and Mexican breeding assemblages. Anim. Behav. 62, 313–329 (2001).Article 

    Google Scholar 
    51.Stimpert, A. K., Peavey, L. E., Friedlaender, A. S. & Nowacek, D. P. Humpback whale song and foraging behavior on an antarctic feeding ground. PLoS ONE 7, e51214 (2012).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    52.Clark, C. W. & Clapham, P. J. Acoustic monitoring on a humpback whale (Megaptera novaeangliae) feeding ground shows continual singing into late spring. Proc. R. Soc. B Biol. Sci. 271, 1051–1057 (2004).Article 

    Google Scholar 
    53.Mattila, D., Guinee, L. & Mato, C. Humpback whale songs on a North Atlantic feeding ground. J. Mammal. 68, 880–883 (1987).Article 

    Google Scholar 
    54.Fournet, M. E. H., Szabo, A. & Mellinger, D. K. Repertoire and classification of non-song calls in southeast Alaskan humpback whales (Megaptera novaeangliae). J. Acoust. Soc. Am. 137, 1–10 (2015).ADS 
    PubMed 
    Article 

    Google Scholar 
    55.Dunlop, R. A., Cato, D. H. & Noad, M. J. Non-song acoustic communication in migrating humpback whales (Megaptera novaeangliae). Mar. Mamm. Sci. 24, 613–629 (2008).Article 

    Google Scholar 
    56.Dunlop, R. A., Noad, M. J., Cato, D. H. & Stokes, D. M. The social vocalization repertoire of east Australian migrating humpback whales (Megaptera novaeangliae). J. Acoust. Soc. Am. 122, 2893–2905 (2007).ADS 
    PubMed 
    Article 

    Google Scholar 
    57.Fournet, M. E. H. et al. Some things never change: Multi-decadal stability in humpback whale calling repertoire on southeast Alaskan foraging grounds. Sci. Rep. 8, 13186 (2018).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    58.Zoidis, A. M. et al. Vocalizations produced by humpback whale (Megaptera novaeangliae) calves recorded in Hawaii. J. Acoust. Soc. Am. 123, 1737–1746 (2008).ADS 
    PubMed 
    Article 

    Google Scholar 
    59.Winn, H. E. et al. Song of the humpback whale: Population comparisons. Behav. Ecol. Sociobiol. 8, 41–46 (1981).Article 

    Google Scholar 
    60.Rekdahl, M. L. et al. Culturally transmitted song exchange between humpback whales (Megaptera novaeangliae) in the southeast Atlantic and southwest Indian ocean basins. R. Soc. Open Sci. 5, 172305 (2018).ADS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    61.Rekdahl, M. L., Dunlop, R. A., Noad, M. J. & Goldizen, A. W. Temporal stability and change in the social call repertoire of migrating humpback whales. J. Acoust. Soc. Am. 133, 1785–1795 (2013).ADS 
    PubMed 
    Article 

    Google Scholar 
    62.Rekdahl, M. L., Tisch, C., Cerchio, S. & Rosenbaum, H. Common nonsong social calls of humpback whales (Megaptera novaeangliae) recorded off northern Angola, southern Africa. Mar. Mamm. Sci. 33, 365–375 (2017).Article 

    Google Scholar 
    63.McDonald, M. A., Calambokidis, J., Teranishi, A. M. & Hildebrand, J. A. The acoustic calls of blue whales off California with gender data. J. Acoust. Soc. Am. 109, 1728–1735 (2002).ADS 
    Article 

    Google Scholar 
    64.Nikolich, K. & Towers, J. R. Vocalizations of common minke whales (Balaenoptera acutorostrata) in an eastern North Pacific feeding ground. Bioacoustics 29, 97–108 (2020).Article 

    Google Scholar 
    65.Delarue, J. Nortwest Atlantic Fin Whale Vocalizations: Geographic Variations and Implications for Stock Assessments (Springer, 2008).
    Google Scholar 
    66.Stimpert, A. K., Au, W. W. L., Parks, S. E., Hurst, T. & Wiley, D. N. Common humpback whale (Megaptera novaeangliae) sound types for passive acoustic monitoring. J. Acoust. Soc. Am. 129, 476–482 (2011).ADS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    67.Dunlop, R. A. Potential motivational information encoded within humpback whale non-song vocal sounds. J. Acoust. Soc. Am. 141, 2204–2213 (2017).ADS 
    PubMed 
    Article 

    Google Scholar 
    68.Wild, L. A. & Gabriele, C. M. Putative contact calls made by humpback whales (Megaptera novaeangliae) in southeastern Alaska. Can. Acoust. 42, 23–31 (2014).
    Google Scholar 
    69.Fournet, M. E. H. Social Calling Behavior of Southeast Alaskan Humpback Whales (Megaptera novaeangliae): Classification and Context (Oregon State University, 2014).
    Google Scholar 
    70.Zerbini, A. N., Clapham, P. J. & Wade, P. R. Assessing plausible rates of population growth in humpback whales from life-history data. Mar. Biol. 157, 1225–1236 (2010).Article 

    Google Scholar 
    71.Gabriele, C. M., Straley, J. M. & Neilson, J. L. Age at first calving of female humpback whales in southeastern Alaska. Mar. Mamm. Sci. 23, 226–239 (2007).Article 

    Google Scholar 
    72.Mizroch, S. A. et al. Estimating the adult survival rate of central North Pacific humpback whales (Megaptera Novaeangliae). J. Mamm. 85, 963–972 (2005).Article 

    Google Scholar 
    73.Whitehead, H. Structure and stability of humpback whale groups off Newfoundland. Can. J. Zool. 61, 1391–1397 (1983).Article 

    Google Scholar 
    74.Tyack, P. L. Functional aspects of cetacean communication. In Cetacean Societies: Field Studies of Dolphins and Whales (ed. Mann, J.) 270–307 (University of Chicago Press, 2000).
    Google Scholar 
    75.Riesch, R., Ford, J. K. B. & Thomsen, F. Stability and group specificity of stereotyped whistles in resident killer whales, Orcinus orca, off British Columbia. Anim. Behav. 71, 79–91 (2006).Article 

    Google Scholar 
    76.Morton, E. S. On the occurrence and significance of motivation-structural rules in some bird and mammal sounds. Am. Nat. 111, 855–869 (1977).Article 

    Google Scholar 
    77.Bradbury, J. W. & Vehrencamp, S. L. Principles of Animal Communication (Springer, 2012).
    Google Scholar 
    78.Wiley, R. H. & Richards, D. G. Physical constraints on acoustic communication in the atmosphere: Implications for the evolution of animal vocalizations. Behav. Ecol. Sociobiol. 3, 69–94 (1978).Article 

    Google Scholar 
    79.Johnson, K. F. & Davoren, G. K. Distributional patterns of humpback whales (Megaptera novaeangliae) along the Newfoundland East Coast reflect their main prey, capelin (Mallotus villosus). Mar. Mamm. Sci. 37, 80–97 (2021).Article 

    Google Scholar 
    80.Rekdahl, M. L. et al. Non-song social call bouts of migrating humpback whales. J. Acoust. Soc. Am. 137, 3042–3053 (2015).ADS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    81.Epp, M. V., Fournet, M. E. H. & Davoren, G. K. Humpback whale call repertoire on a northeastern Newfoundland foraging ground. Mar. Mamm. Sci. 1–18. https://doi.org/10.1111/mms.12859 (2021).82.Rossi-santos, M. R. Oil industry and noise pollution in the humpback whale (Megaptera novaeangliae) soundscape ecology of the Southwestern Atlantic breeding ground. J. Coast. Res. 31, 184–195 (2015).Article 

    Google Scholar 
    83.Cholewiak, D. M. et al. Communicating amidst the noise: Modeling the aggregate influence of ambient and vessel noise on baleen whale communication space in a national marine sanctuary. Endanger. Species Res. 36, 59–75 (2018).Article 

    Google Scholar 
    84.Bioacoustics Research Program. Raven Pro: Interactive Sound Analysis Software (Version 1.5) [Computer Software]. (2014).85.Mellinger, D. K. & Bradbury, J. W. Acoustic measurement of marine mammal sounds in noisy environments. In Proceedings of the Second International Conference on Underwater Acoustic Measurements: Technologies and Results, Heraklion, Greece 8 (2007).86.Epp, M. V. The Call Repertoire of Humpback Whales (Megaptera novaeangliae) on a Newfoundland Foraging Ground (2015, 2016) with Comparison to a Hawaiian Breeding Ground (1981, 1982) (University of Manitoba, 2019).
    Google Scholar 
    87.Breiman, L., Friedman, J. H., Olshen, R. A. & Stone, C. J. Classification and Regression Trees (Wadsworth International Group, 1984).MATH 

    Google Scholar 
    88.Silber, G. K. Non-song Phonations and Associated Surface Behavior Of the Hawaiian Humpback Whales (San Jose State University, 1986).
    Google Scholar  More

  • in

    Intra- and interspecific variability among congeneric Pagellus otoliths

    Intra- and interspecific differences: comparison with former studies on Pagellus species and other fish speciesTo understand the relationship between function, shape, and the environment, it is essential to include the morphological variability of otoliths, considering biological and environmental variability leads to otolith shape heterogeneity through morpho-functional adaptation to different habitats. Several authors have highlighted changes in otolith shape between species and, in many cases, among populations of the same species (e.g., herrings, salmonids, and lutjanids). The intra-specific variability of otolith morphology and shape are the basis of stock separation and assessment and is related, especially in sagittae, with environmental (e.g., water temperature, salinity, and depth) and biological factors (e.g., sex, ontogeny, and genetic variability)20.The analysis of the three Pagellus species revealed that otolith morphology and morphometry did not follow those described in a previous study1 conducted in the western Mediterranean Sea and the Atlantic Ocean in term of rectangularity, circularity, sagitta aspect ratio and sagitta length to total fish length ratio. Although the images provided in our study closely resembled those from research in other geographical areas, the morphometric measures (obtained according to the procedures and methods described in the previous literature1,20,23,32) exhibited several differences. Considering the scale of our study compared to previous studies, it is difficult to provide an entirely valid comparison; the differences in sagitta morphology and morphometry could have been triggered by biotic and abiotic parameters (e.g., temperature, salinity, genotype, habitat type, differences in food quality and quantity)13,33,34,35. Such environmental and genetic factors may be primary drivers of otolith morphometry and morphology among fishes in different habitats. Therefore, detected shape differences are at the basis of fish stock differentiation36.Our results indicate that the min–max circularity and rectangularity of P. bogaraveo from the Southern Tyrrhenian Sea differ from those calculated in a previous study1 in the western Mediterranean Sea, and the north and central-eastern Atlantic ocean. Moreover, the increase in circularity in larger specimens, confirms a greater tendency toward circular than elliptical otolith shape in southern Tyrrhenian Sea species compared to those in other Mediterranean and Atlantic areas.Despite statistical differences and correlations in this study supported the hypothesis that some changes in sagitta morphology are related to fish size differences, several aspects and studies should be performed to better understand this relation. The negative correlation between the ratio of sulcus acusticus surface to the entire sagitta, rostral morphology, and the increase in specimen’s size was related to the expansion in the length and surface of the entire sagitta and rostral area in larger specimens. These features, with no statistical relevant increment in sulcus acusticus surface and increased rostrum length, could be correlated with more pronounced peripheral sagitta growth in this species. Sagitta, in fact, after fish pelagic phase, might increases its surface in the rostral area and the margins. Since the present study did not take into account ontogenetic stages and specimens age, it is hard to relate this result with sagitta and sulcus acusticus growth. But reading this increase by an ecological point of view, it could be related to the lifecycle of the species. During the juvenile stage, in the early stage of pelagic life, the species inhabits shallow water. Adults inhabit deep-water environments, migrating down the continental slope to a depth of 800 m after the juvenile stage. These changes in habitat might be the cause of morphological variations in the sagittae, highlighting the relationship between sagitta features and environmental and biological factors.The P. acarne specimens demonstrated the highest number of morphometrical parameters that did not follow those of the same species described in a previous study (i.e., circularity, rectangularity, sagitta length to total fish length ratio, and sagitta aspect ratio)1. These morphometrical changes are reflected in otolith shape. The otoliths from specimens in our study were largely circular, with highly irregular margins and a rostrum that varied in length and width through the left and right sagitta, as indicated by the significant differences in rostrum aspect ratio values.The morphometrical results in P. erythrinus revealed differences in circularity, rectangularity, sagitta length to total fish length ratio and sagitta aspect ratio compared to a previous study1 in the western Mediterranean Sea and north-central eastern Atlantic ocean.The P. erythrinus specimens were characterized by a pentagonal otoliths shape, and increased circularity compared to the same species from other areas. The results also indicated small differences between the left and right sagitta. This small differences were previously described in other Mediterranean sub-areas, for example, otolith width values in P. erythrinus specimens collected in the Gulf of Tunisia37,38.As said above for P. bogaraveo, it is hard to relate the differences between juveniles and adults with fish growth due to the absence in present paper of ontogenetic and age analysis. The higher width than length, demonstrated by min–max width values in Tables 1 and 2, in sagittae of adults P. erythrinus specimens could be correlated with an exponential increment in fish size compared to the sagittal length. Further analyses on ontogenetic development of this species are required to better define the sagitta growth related to fish growth.The increase in sulcus acusticus surface exhibited in the adult specimens could be correlated with feeding habits; during its adult life, this species is a benthic feeder and inhabits deeper environments than juveniles28,29.Although meaningful lateral dimorphism of the sagittae was detected only in flatfish, statistical analysis revealed several small differences between the left and right sagitta in P. erythrinus and P. acarne, as previously described in other round fish species, such as Chelon ramada (Risso, 1827)40, Diplodus annularis (Linnaeus, 1758)41, Diplodus puntazzo (Walbaum, 1792)42, Clupea harengus (Linnaeus, 1758)43, and Scomberomorus niphonius (Cuvier, 1832)44.Our study confirmed slight differences between width values in left and right sagitta previously described in P. erythrinus and extend the differences to other parameters, such as circularity and rectangularity (Tables 1, 2). Concerning P. acarne, however, marginal differences between the left and right sagittae were observed for the first time.This slight differences are supported by the literature concerning genetic and environmental stressors41. Since the functional morphology of otoliths is not completely understood, it is difficult to find a direct link between these small differences and the ecology of the species. However, several eco-functional factors, such as feeding behavior, deserve attention as fundamental for a better understanding of the relationship between otolith features and species habitat. For example, P. erythrinus largely preys on strictly benthic organisms, such as polychaetes, brachyuran crabs, and benthic crustaceans. Most of these species frequently escape predators by hiding under the sandy substrate. Other Sparidae (Lythognathus mormyrus, Linnaeus, 1758) feed on benthic fauna, engulfing sediment and filtering it in the buccal cavity, demonstrated by the high percentage of detritus and benthic remains (e.g., scales, urchin spines, and benthic foraminifers) in the gut and stomach contents29. To engulf sediment, P. erythrinus performs a particular movement with the head and body, laterally shifting and pushing forward, to dig the bottom sand and reach prey. This kind of behavior, common in all benthopelagic species with the same feeding habits, could influence the sagitta growth and morphology, triggering small differences between the left and right sagitta. Further studies on this and other species with this behavior (e.g., L. mormyrus) are necessary to confirm this hypothesis.Concerning inter-specific differences in sagitta morphology among the three species, it is difficult to read the results obtained in this study eco-morphologically since an insufficient understanding of the functional morphology and physiology of otoliths prohibits a direct relationship, valid for all the species, between eco-functional features and otolith morphology. Nevertheless, as expected, the shape analysis (Fig. 1) revealed clear differences between the three congeneric species. Considering several ecological, functional, and biological features in each species, the results have demonstrated a sagitta morphology that could be in accordance with the ecology and lifestyle of these three congeneric seabreams.Relationship between otolith morphology and ecology/lifestyleThe sagittae of P. acarne exhibited a shape resembling those in other pelagic species, with a long rostrum and the entire sagitta elongated and narrower than those in other two seabream species. The species that show the most pelagic habits, with largely planktivorous feeding at a small size, adapt also to benthopelagic feeding activity in adult life. The statistically relevant similarity found in P. bogaraveo could be proof of the ecomorphological adaptation of sagittae to pelagic and demersal environments. This hypothesis may be confirmed by marked differences in shape compared to those in P. erythrinus, which is the most benthic among the three species.Pagellus erythrinus was the species with the shortest rostrum. It also has the most benthic habits, largely preying on epibenthic and infaunal species. Moreover, its ecology and life cycle differ among the three species under study since they are strictly related to the benthic environment. This lifestyle could be in accordance with the differences observed in the shape analysis results. The sagitta contours appeared more circular and wider than those in the other two species. The PCA and LDA also confirmed the most difference in shape among the three species.The species with the most marked antirostrum and sagitta shape was P. bogaraveo, which is a cross between the other two congeneric species. Pagellus bogaraveo is a demersal species, which inhabits the deep biocenosis and feeds in both benthic and mesopelagic environments. Furthermore, the ecology of this species could support the sagitta shape described in our study27,30.Otolith morphology and morphometry in congeneric Pagellus species described in this study has followed the relationship between sagittal parameters, habitat, and depth described in previous literature15. According to several authors, the percentage of species with large otoliths increases with depth, except for abyssal depth. The specimens of P. bogaraveo analyzed in this paper (especially adult individuals) had larger otoliths than the other two Pagellus species due to their demersal habits (they inhabit the continental slope to a depth of 800 m). A larger sagitta is essential in demersal environments to compensate for light reduction by providing improved acoustic communication, sound perception15,45, and a sense of equilibrium46.Sulcus shapeConsidering the sulcus acusticus, in the otolith atlas for the western Mediterranean Sea and Atlantic ocean1, studies describing and comparing otoliths10 and the diversity and variability of otoliths in teleost fishes9, the sulcus in P. bogaraveo, P. acarne and P. erythrinus was described as heterosulcoid, with an ostium shorter than the cauda and a long, narrowed rostrum, especially in adult P. bogaraveo and P. acarne individuals. Heterosulcoid otoliths were also observed in south Tyrrhenian Sea Pagellus individuals, with marked differences between juvenile and adult specimens. In a demersal species, such as P. bogaraveo, juveniles live in shallow, coastal water. Once adults, they inhabit deeper water (to a depth of 800 m). Changes in the crystalline and morphological structure of sulcus acusticus between juveniles and adults reflect this species’ need to adapt to deeper environments with less light.The results indicate that in P. bogaraveo, the sulcus acusticus does not differ in surface between juvenile and adult specimens. This feature could be correlated with earlier sulcus acusticus development in this species, compared to P. erythrinus and P. acarne, emphasizing the role of the sulcus acusticus in this demersal species48,49. This might also confirm the strict correlation between biological and environmental factors and sagitta morphology in studied seabreams species.Another morphological feature of the sulcus acusticus, which might support the ecology of the species, is the deep ostium and cauda. In adult specimens of P. bogaraveo and P. acarne, the sulcus structure deeply penetrated in the sagitta carbonate structure. Conversely, in adult P. erythrinus specimens, the sulcus did not penetrate as deeply as in the other two seabream species. This sulcal feature could correspond with the ecology and feeding behavior of P. erythrinus, which specializes in benthic strategies, including small differences between left and right sagitta and the absence of the notch and antirostrum in sagittae.Although the deeper sulcus acusticus in P. bogaraveo and P. acarne might be linked to depth distribution, as in P. bogaraveo, it may also correspond with high mobility related to feeding behavior, as in both P. bogaraveo and P. acarne. The different depths of sulcus acusticus can change the thickness of the otolithic membrane, by varying the relative motion of otoliths with the macula sacculi2. As previously demonstrated50, the different thicknesses of the otolithic membrane induce differences in mechanical resistance between the otolith and sensory epithelium.The differences in sulcus acusticus and otolith ratio between P. bogaraveo specimens and the other congeneric species, demonstrated by the results, might be also correlated to the differences in habitat, feeding habits, and soundscape.Despite the lack of information concerning the physiological ear response related to variations in macula or sulcus size, the sensory hair cells in macula sacculi are likely to be affected by changes in sulcus depth, shape, 3D structure (planar vs. curved), and surface.The significant difference in relative sulcus area may be due to typical alteration in this parameter concerning differences in the mobility patterns, food, feeding behavior, and spatial niche.Higher relative sulcus area ratios have been observed in the deepest species or those with high mobility49. In our study, the morphometry results concerning the sulcus did not follow those in the previous literature, displaying higher values in P. erythrinus and P. acarne compared to P. bogaraveo, although the latter inhabits a deeper environment than the other congeneric species.This higher relative sulcus acusticus surface and the larger, curved sulcus acusticus of P. acarne and P. erythrinus could be correlated with higher mobility in these species (especially P. acarne). In P. erythrinus, however, these features might be related to its benthic lifestyle.As demonstrated by the PCA and LDA of sulcus acusticus parameters, P. erythrinus and P. acarne, which share similar depths and habitats, revealed marked similarities, whereas P. bogaraveo, which lives in the deepest strata of the water column, displayed the most different sulcus acusticus. However, PCA and LDA indicated that the otolith shape in the entire P. erythrinus sagitta was significantly different compared to those in P. acarne and P. bogaraveo.These features could provide a reading key for sagitta and sulcus acusticus eco-morphology in the life cycle and environmental adaptation of fish.The connection between the otoliths and the macula sacculi is fundamental for transducing environmental acoustic signals and for the relative motion of fish (balance). The sulcus acusticus is the area of the otoliths in which this connection occurs.Features of the textureFurthermore, the external textural organization23 changes between juveniles and adults or when environmental changes occur. The differences in the external textural organization found in juveniles and adults support those reported in the literature concerning other species39. Figures 3b, c and 5a–h, demonstrate that our study supported this prediction. However, P. bogaraveo and P. erythrinus juveniles, compared with other species (such as gurnards)23 displayed a more uniform, mineralized, external textural organization.Figure 5SEM images of the crystalline structure of P. bogaraveo, juveniles (a, b) and adults (c, d), and P. erythrinus, juveniles (e, f) and adults (g, h) sagittae.Full size imageAccording to previous literature8, improved hearing capabilities in a species are closely related to a higher value of relative sulcus area ratio. Habitat features, such as depth, feeding strategies, mobility, trophic distribution, and ontogeny, could also influence this ratio.Hence, it may be concluded that morphological differences in sulcus acusticus shape and surface among species are important for comprehending the ecomorphological and eco-functional role of sagitta 2,48.Comparing the intra-specific differences indicated by our results with those in the literature, discussing other populations, we cannot determine whether site-differences observed in sagitta shape are related to genetic evolution and/or adaptative response to environment. To make this distinction it would require a specific experiment in which offspring from different populations are raised in a controlled environment.Furthermore, the knowledge about physiology and functional morphology is insufficient to provide a clear correlation between inter-specific differences among the three congeneric Pagellus species and their ecological and functional features. However, differences in sagitta morphology and morphometry among these three Pagellus species may be related to differences in lifestyle, ecology, and biology since they follow the ecomorphological features of sagittae and species ecology described in the literature. More