More stories

  • in

    Organic carbon accumulation and aggregate formation in soils under organic and inorganic fertilizer management practices in a rice–wheat cropping system

    Zhe, W. et al. Probing the nature of soil organic matter. Crit. Rev. Environ. Sci. Technol. 52, 4072–4093 (2022).Article 

    Google Scholar 
    Blanco-Canqui, H. & Lal, R. Mechanisms of carbon sequestration in soil aggregates. Crit. Rev. Plant Sci. 23, 481–504 (2004).Article 
    CAS 

    Google Scholar 
    Six, J., Paustian, K., Elliott, E. T. & Combrink, C. Soil structure and organic matter: I. Distribution of aggregate–size classes and aggregate–associated carbon. Soil Sci. Soc. Am. J. 64, 681–689 (2000).Article 
    ADS 
    CAS 

    Google Scholar 
    Lehmann, J. & Kleber, M. The contentious nature of soil organic matter. Nature 528, 60–68 (2015).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Six, J., Bossuyt, H., Degryze, S. & Denef, K. A history of research on the link between (micro)aggregates, soil biota, and soil organic matter dynamics. Soil Tillage Res. 79, 7–31 (2004).Article 

    Google Scholar 
    Tisdall, J. M. & Oades, J. M. Organic matter and water-stable aggregates in soils. Eur. J. Soil Sci. 33, 141–163 (1982).Article 
    CAS 

    Google Scholar 
    Luo, Y. et al. Rice rhizodeposition promotes the build-up of organic carbon in soil via fungal necromass. Soil Biol. Biochem. 160, 108345 (2021).Article 
    CAS 

    Google Scholar 
    Wang, X. et al. Organic amendments drive shifts in microbial community structure and keystone taxa which increase C mineralization across aggregate size classes. Soil Biol. Biochem. 153, 108062 (2021).Article 
    CAS 

    Google Scholar 
    Duan, Y. et al. Long–term fertilisation reveals close associations between soil organic carbon composition and microbial traits at aggregate scales. Agric. Ecosyst. Environ. 306, 107169 (2021).Article 
    CAS 

    Google Scholar 
    Christensen, B. T. Physical fractionation of soil and structural and functional complexity in organic matter turnover. Eur. J. Soil Sci. 52, 345–353 (2001).Article 
    CAS 

    Google Scholar 
    Olk, D. C. & Gregorich, E. G. Overview of the symposium proceedings, “meaningful pools in determining soil carbon and nitrogen dynamics”. Soil Sci. Soc. Am. J. 70, 967–974 (2006).Article 
    ADS 
    CAS 

    Google Scholar 
    Courtier-Murias, D. et al. Unraveling the long–term stabilization mechanisms of organic materials in soils by physical fractionation and NMR spectroscopy. Agric. Ecosyst. Environ. 171, 9–18 (2013).Article 
    CAS 

    Google Scholar 
    Rodrigues, L. A. T. et al. Short– and long–term effects of animal manures and mineral fertilizer on carbon stocks in subtropical soil under no–tillage. Geoderma 386, 114913 (2021).Article 
    ADS 
    CAS 

    Google Scholar 
    Mao, J., Dan, C. O., Fang, X., He, Z. & Schmidt-Rohr, K. Influence of animal manure application on the chemical structures of soil organic matter as investigated by advanced solid–state NMR and FT–IR spectroscopy. Geoderma 146, 353–362 (2008).Article 
    ADS 
    CAS 

    Google Scholar 
    Simonetti, G. et al. Characterization of humic carbon in soil aggregates in a long–term experiment with manure and mineral fertilization. Soil Sci. Soc. Am. J. 25, 880–890 (2012).Article 

    Google Scholar 
    Cambardella, C. A. & Elliott, E. T. Particulate soil organic-matter changes across a grassland cultivation sequence. Soil Sci. Soc. Am. J. 56, 777–783 (1992).Article 
    ADS 

    Google Scholar 
    Conant, R. T., Six, J. & Paustian, K. Land use effects on soil carbon fractions in the southeastern United States. I. Management-intensive versus extensive grazing. Biol. Fertil. Soils 38, 386–392 (2003).Article 
    CAS 

    Google Scholar 
    Blanco-Moure, N., Gracia, R., Bielsa, A. C. & López, M. V. Soil organic matter fractions as affected by tillage and soil texture under semiarid Mediterranean conditions. Soil Tillage Res. 155, 381–389 (2016).Article 

    Google Scholar 
    Yu, H. et al. Accumulation of organic C components in soil and aggregates. Sci. Rep. 5, 13804 (2015).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Schöning, I., Morgenroth, G. & Kögel-Knabner, I. O/N–alkyl and alkyl C are stabilised in fine particle size fractions of forest soils. Biogeochemistry 73, 475–497 (2005).Article 

    Google Scholar 
    Solomon, D., Lehmann, J., Kinyangi, J., Liang, B. & Schäfer, T. Carbon K-edge NEXAFS and FTIR–ATR spectroscopic investigation of organic carbon speciation in soils. Soil Sci. Soc. Am. J. 13, 107–119 (2005).Article 

    Google Scholar 
    Yan, H., Chen, C., Xu, Z., Williams, D. & Xu, J. Assessing management impacts on soil organic matter quality in subtropical Australian forests using physical and chemical fractionation as well as 13C NMR spectroscopy. Soil Biol. Biochem. 41, 640–650 (2009).Article 

    Google Scholar 
    Masoom, H. et al. Soil organic matter in its native state: Unravelling the most complex biomaterial on earth. Environ. Sci. Technol. 50, 1670–1680 (2016).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Vogel, C. et al. Clay mineral composition modifies decomposition and sequestration of organic carbon and nitrogen in fine soil fractions. Biol. Fertil. Soils 51, 427–442 (2015).Article 
    CAS 

    Google Scholar 
    Sharma, S., Singh, P., Angmo, P. & Satpute, S. Total and labile pools of organic carbon in relation to soil biological properties under contrasting land-use systems in a dry mountainous region. Carbon Manage. 13, 352–371 (2022).Article 
    CAS 

    Google Scholar 
    Six, J., Elliott, E., Paustian, K. & Doran, J. Aggregation and soil organic matter accumulation in cultivated and native grassland soils. Soil Sci. Soc. Am. J. 62, 1367–1377 (1998).Article 
    ADS 
    CAS 

    Google Scholar 
    Elliott, E. T. Aggregate structure and carbon, nitrogen, and phosphorus in native and cultivated soils. Soil Sci. Soc. Am. J. 50, 627–633 (1986).Article 
    ADS 

    Google Scholar 
    Yu, H., Ding, W., Luo, J., Geng, R. & Cai, Z. Long-term application of organic manure and mineral fertilizers on aggregation and aggregate-associated carbon in a sandy loam soil. Soil Tillage Res. 124, 170–177 (2012).Article 

    Google Scholar 
    Carter, M. R. & Gregorich, E. G. (eds) Soil Sampling and Methods of Analysis 2nd edn, 230–233 (Taylor & Francis Group, CRC, 2007).
    Google Scholar 
    Lu, R. (ed.) Soil and Agro-chemistry Analytical Methods 146–149 (China Agricultural Science and Technology Press, 1999).
    Google Scholar 
    Wu, J., Joergensen, R. G., Pommerening, B., Chaussod, R. & Brookes, P. C. Measurement of soil microbial biomass C by fumigation extraction: An automated procedure. Soil Biol. Biochem. 22, 1167–1169 (1990).Article 
    CAS 

    Google Scholar 
    Zhang, X., Zhu, A., Yang, W. & Zhang, J. Accumulation of organic components and its association with macroaggregation in a sandy loam soil following conservation tillage. Plant Soil. 416, 1–15 (2017).Article 
    CAS 

    Google Scholar 
    Skjemstad, J. O., Clarke, P., Taylor, J. A., Oades, J. M. & Newman, R. H. The removal of magnetic materials from surface soils—a solid state 13C CP/MAS NMR study. Soil Res. 32, 1215–1229 (1994).Article 
    CAS 

    Google Scholar 
    Ringle, C. M., Wende, S. & Becker, J. M. SmartPLS 3.” Boenningstedt: SmartPLS GmbH. Preprint at http://www.smartpls.com (2015).Jerbi, M., Labidi, S., Lounès-Hadj Sahraoui, A., Chaar, H. & Ben Jeddi, F. Higher temperatures and lower annual rainfall do not restrict, directly or indirectly, the mycorrhizal colonization of barley (Hordeum vulgare L.) under rainfed conditions. PLoS ONE 15, e0241794 (2020).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Cohen, J. Statistical power analysis for the behavioral sciences 2nd edn, 407–530 (Erlbaum Associates, Berlin, 1988).MATH 

    Google Scholar 
    Singh, P. & Benbi, D. K. Physical and chemical stabilization of soil organic matter in cropland ecosystems under rice–wheat, maize–wheat and cotton–wheat cropping systems in northwestern India. Carbon Manag. 12, 603–621 (2021).Article 
    CAS 

    Google Scholar 
    Kiem, R. & Kögel-Knabner, I. Contribution of lignin and polysaccharides to the refractory carbon pool in C–depleted arable soils. Soil Biol. Biochem. 35, 101–118 (2003).Article 
    CAS 

    Google Scholar 
    Lutzow, M. V. et al. Stabilization of organic matter in temperate soils: mechanisms and their relevance under different soil conditions: a review. Eur. J. Soil Sci. 57, 426–445 (2006).Article 

    Google Scholar 
    Yudina, A. V., Klyueva, V. V., Romanenko, K. A. & Fomin, D. S. Micro- within macro: How micro-aggregation shapes the soil pore space and water-stability. Geoderma 415, 115771 (2022).Article 
    ADS 
    CAS 

    Google Scholar 
    Tisdall, J. M., Smith, S. E. & Rengasamy, P. Aggregation of soil by fungal hyphae. Soil Res. 35, 55–60 (1997).Article 

    Google Scholar 
    Li, T. et al. Contrasting impacts of manure and inorganic fertilizer applications for nine years on soil organic carbon and its labile fractions in bulk soil and soil aggregates. CATENA 194, 104739 (2020).Article 
    CAS 

    Google Scholar 
    Liang, Y. et al. Effect of chemical fertilizer and straw-derived organic amendments on continuous maize yield, soil carbon sequestration and soil quality in a Chinese Mollisol. Agric. Ecosyst. Environ. 314, 107403 (2021).Article 
    CAS 

    Google Scholar 
    Liang, C., Kästner, M. & Joergensen, R. G. Microbial necromass on the rise: The growing focus on its role in soil organic matter development. Soil Biol. Biochem. 150, 108000 (2020).Article 
    CAS 

    Google Scholar 
    Sharma, S., Singh, P. & Kumar, S. Responses of soil carbon pools, enzymatic activity, and crop yields to nitrogen and straw incorporation in a rice-wheat cropping system in North-Western India. Front. Sustain. Food Syst. 4, 532704 (2020).Article 

    Google Scholar 
    Puget, P., Chenu, C. & Balesdent, J. Dynamics of soil organic matter associated with particle–size fractions of water–stable aggregates. Eur. J. Soil Sci. 51, 595–605 (2000).Article 

    Google Scholar  More

  • in

    Habitat partitioning, co-occurrence patterns, and mixed-species group formation in sympatric delphinids

    Pianka, E. R. Niche overlap and diffuse competition. Proc. Natl. Acad. Sci. 71, 2141–2145 (1974).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Chesson, P. Mechanisms of maintenance of species diversity. Annu. Rev. Ecol. Syst. 31, 343–366 (2000).Article 

    Google Scholar 
    Tokeshi, M. Species Coexistence: Ecological and Evolutionary Perspectives. (Wiley-Blackwell, 2009).Grinnell, J. Geography and evolution. Ecology 5, 225–229 (1924).Article 

    Google Scholar 
    Roughgarden, J. Resource partitioning among competing species—A coevolutionary approach. Theor. Popul. Biol. 9, 388–424 (1976).Article 
    MathSciNet 
    CAS 
    PubMed 
    MATH 

    Google Scholar 
    Syme, J., Kiszka, J. J. & Parra, G. J. Dynamics of cetacean mixed-species groups: A review and conceptual framework for assessing their functional significance. Front. Mar. Sci. 8, 1–19 (2021).Article 

    Google Scholar 
    Stensland, E., Angerbjörn, A. & Berggren, P. Mixed species groups in mammals. Mamm. Rev. 33, 205–223 (2003).Article 

    Google Scholar 
    Cords, M. & Würsig, B. A Mix of Species: Associations of Heterospecifics Among Primates and Dolphins. in Primates and Cetaceans: Field Research and Conservation of Complex Mammalian Societies (eds. Yamagiwa, J. & Karczmarski, L.) 409–431 (Springer, 2014). doi:https://doi.org/10.1007/978-4-431-54523-1_21.Goodale, E., Beauchamp, G. & Ruxton, G. D. Mixed-Species Groups of Animals: Behavior, Community Structure, and Conservation. (Academic Press, 2017).Krause, J. & Ruxton, G. D. Living in Groups. Oxford Series in Ecology and Evolution (Oxford University Press, 2002).Heymann, E. W. & Buchanan-Smith, H. M. The behavioural ecology of mixed-species troops of callitrichine primates. Biol. Rev. 75, 169–190 (2000).Article 
    CAS 
    PubMed 

    Google Scholar 
    Sridhar, H. & Guttal, V. Friendship across species borders: factors that facilitate and constrain heterospecific sociality. Philos. Trans. R. Soc. B Biol. Sci. 373, 1–9 (2018).Greenberg, R. Birds of many feathers: The formation and structure of mixed-species flocks of forest birds. in On the Move: How and Why Animals Travel in groups (eds. Boinski, S. & Gerber, P. A.) 521–558 (University of Chicago Press, 2000).Waser, P. M. ‘Chance’ and mixed-species associations. Behav. Ecol. Sociobiol. 15, 197–202 (1984).Article 

    Google Scholar 
    Whitesides, G. H. Interspecific associations of Diana monkeys, Cercopithecus diana, in Sierra Leone, West Africa: biological significance or chance?. Anim. Behav. 37, 760–776 (1989).Article 

    Google Scholar 
    Waser, P. M. Primate polyspecific associations: Do they occur by chance?. Anim. Behav. 30, 1–8 (1982).Article 

    Google Scholar 
    Alexander, R. D. The evolution of social behavior. Annu. Rev. Ecol. Syst. 5, 325–383 (1974).Article 

    Google Scholar 
    Kasozi, H. & Montgomery, R. A. Variability in the estimation of ungulate group sizes complicates ecological inference. Ecol. Evol. 10, 6881–6889 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Syme, J., Kiszka, J. J. & Parra, G. J. How to define a dolphin ‘group’? Need for consistency and justification based on objective criteria. Ecol. Evol. 12, 1–18 (2022).Article 

    Google Scholar 
    Hutchinson, J. M. C. & Waser, P. M. Use, misuse and extensions of ‘ideal gas’ models of animal encounter. Biol. Rev. 82, 335–359 (2007).Article 
    PubMed 

    Google Scholar 
    Gotelli, N. J. Null model analysis of species co-occurrence patterns. Ecology 81, 2606–2621 (2000).Article 

    Google Scholar 
    Astaras, C., Krause, S., Mattner, L., Rehse, C. & Waltert, M. Associations between the drill (Mandrillus leucophaeus) and sympatric monkeys in Korup National Park. Cameroon. Am. J. Primatol. 73, 127–134 (2011).Article 
    PubMed 

    Google Scholar 
    Mammides, C., Chen, J., Goodale, U. M., Kotagama, S. W. & Goodale, E. Measurement of species associations in mixed-species bird flocks across environmental and human disturbance gradients. Ecosphere 9, 1–14 (2018).Article 

    Google Scholar 
    Ovaskainen, O., Abrego, N., Halme, P. & Dunson, D. Using latent variable models to identify large networks of species-to-species associations at different spatial scales. Methods Ecol. Evol. 7, 549–555 (2016).Article 

    Google Scholar 
    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 
    Warton, D. I. et al. So Many variables: Joint modeling in community ecology. Trends Ecol. Evol. 30, 766–779 (2015).Article 
    PubMed 

    Google Scholar 
    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).Article 
    PubMed 

    Google Scholar 
    Ovaskainen, O. & Abrego, N. Joint Species Distribution Modelling. (Cambridge University Press, 2020). https://doi.org/10.1017/9781108591720.Blanchet, F. G., Cazelles, K. & Gravel, D. Co-occurrence is not evidence of ecological interactions. Ecol. Lett. 23, 1050–1063 (2020).Article 
    PubMed 

    Google Scholar 
    Haak, C. R., Hui, F. K., Cowles, G. W. & Danylchuk, A. J. Positive interspecific associations consistent with social information use shape juvenile fish assemblages. Ecology 101, 1–16 (2020).Article 

    Google Scholar 
    Bastianelli, G., Wintle, B. A., Martin, E. H., Seoane, J. & Laiolo, P. Species partitioning in a temperate mountain chain: Segregation by habitat vs. interspecific competition. Ecol. Evol. 7, 2685–2696 (2017).Aspin, T. & House, A. Alpha and beta diversity and species co-occurrence patterns in headwaters supporting rare intermittent-stream specialists. Freshw. Biol. n/a, (2022).Astarloa, A. et al. Identifying main interactions in marine predator-prey networks of the Bay of Biscay. ICES J. Mar. Sci. 76, 2247–2259 (2019).Article 

    Google Scholar 
    Parra, G. J. Resource partitioning in sympatric delphinids: space use and habitat preferences of Australian snubfin and Indo-Pacific humpback dolphins. J. Anim. Ecol. 75, 862–874 (2006).Article 
    PubMed 

    Google Scholar 
    Parra, G. J., Wojtkowiak, Z., Peters, K. J. & Cagnazzi, D. Isotopic niche overlap between sympatric Australian snubfin and humpback dolphins. Ecol. Evol. 12, 1–11 (2022).Article 

    Google Scholar 
    Kiszka, J. J. et al. Ecological niche segregation within a community of sympatric dolphins around a tropical island. Mar. Ecol. Prog. Ser. 433, 273–288 (2011).Article 
    ADS 

    Google Scholar 
    Bearzi, M. Dolphin sympatric ecology. Mar. Biol. Res. 1, 165–175 (2005).Article 

    Google Scholar 
    Zaeschmar, J. R. et al. Occurrence of false killer whales (Pseudorca crassidens) and their association with common bottlenose dolphins (Tursiops truncatus) off northeastern New Zealand. Mar. Mammal Sci. 30, 594–608 (2014).Article 

    Google Scholar 
    Elliser, C. R. & Herzing, D. L. Long-term interspecies association patterns of Atlantic bottlenose dolphins, Tursiops truncatus, and Atlantic spotted dolphins, Stenella frontalis, in the Bahamas. Mar. Mammal Sci. 32, 38–56 (2016).Article 

    Google Scholar 
    Kiszka, J. J., Perrin, W. F., Pusineri, C. & Ridoux, V. What drives island-associated tropical dolphins to form mixed-species associations in the southwest Indian Ocean?. J. Mammal. 92, 1105–1111 (2011).Article 

    Google Scholar 
    Brown, A. M., Bejder, L., Cagnazzi, D., Parra, G. J. & Allen, S. J. The north west cape, Western Australia: A potential hotspot for Indo-Pacific humpback dolphins Sousa chinensis?. Pacific Conserv. Biol. 18, 240–246 (2012).Article 

    Google Scholar 
    Allen, S. J., Cagnazzi, D., Hodgson, A. J., Loneragan, N. R. & Bejder, L. Tropical inshore dolphins of north-western Australia: Unknown populations in a rapidly changing region. Pacific Conserv. Biol. 18, 56–63 (2012).Article 

    Google Scholar 
    Palmer, C., Parra, G. J., Rogers, T. & Woinarski, J. Collation and review of sightings and distribution of three coastal dolphin species in waters of the Northern Territory. Australia. Pacific Conserv. Biol. 20, 116–125 (2014).Article 

    Google Scholar 
    Corkeron, P. J. Aspects of the Behavioral Ecology of Inshore Dolphins Tursiops truncatus and Sousa chinensis in Moreton Bay, Australia. in The Bottlenose Dolphin (eds. Leatherwood, S. & Reeves, R.) 285–293 (Elsevier, 1990). https://doi.org/10.1016/B978-0-12-440280-5.50018-4.Haughey, R. et al. Distribution and habitat preferences of Indo-Pacific Bottlenose Dolphins (Tursiops aduncus) inhabiting coastal waters with mixed levels of protection. Front. Mar. Sci. 8, 1–20 (2021).Article 

    Google Scholar 
    Hanf, D., Hodgson, A. J., Kobryn, H., Bejder, L. & Smith, J. N. Dolphin distribution and habitat suitability in North Western Australia: Applications and Implications of a Broad-Scale, Non-targeted Dataset. Front. Mar. Sci. 8, 1–18 (2022).Article 

    Google Scholar 
    Hunt, T. N., Allen, S. J., Bejder, L. & Parra, G. J. Identifying priority habitat for conservation and management of Australian humpback dolphins within a marine protected area. Sci. Rep. 10, 1–14 (2020).Article 

    Google Scholar 
    Hunt, T. N. Demography, habitat use and social structure of Australian humpback dolphins (Sousa sahulensis) around the North West Cape, Western Australia: Implications for conservation and management. PhD Thesis, Flinders University, Adelaide, Australia. (Flinders University, 2018).Cassata, L. & Collins, L. B. Coral reef communities, habitats, and substrates in and near sanctuary zones of Ningaloo marine park. J. Coast. Res. 241, 139–151 (2008).Article 

    Google Scholar 
    CALM MPRA. Management plan for the Ningaloo Marine Park and Muiron Islands Marine Management Area 2005–2015. (2005).Hunt, T. N. et al. Demographic characteristics of Australian humpback dolphins reveal important habitat toward the southwestern limit of their range. Endanger. Species Res. 32, 71–88 (2017).Article 

    Google Scholar 
    Mann, J. Behavioral sampling methods for cetaceans: A review and critique. Mar. Mammal Sci. 15, 102–122 (1999).Article 

    Google Scholar 
    Python Software Foundation. Python Language Reference, version 3.8.0. at https://www.python.org/ (2016).QGIS Development Team. QGIS Geographic Information System, version 3.8.3 Zanzibar. at http://qgis.osgeo.org (2019).Zanardo, N., Parra, G., Passadore, C. & Möller, L. Ensemble modelling of southern Australian bottlenose dolphin Tursiops sp. distribution reveals important habitats and their potential ecological function. Mar. Ecol. Prog. Ser. 569, 253–266 (2017).Hanberry, B. B. Finer grain size increases effects of error and changes influence of environmental predictors on species distribution models. Ecol. Inform. 15, 8–13 (2013).Article 

    Google Scholar 
    Gottschalk, T. K., Aue, B., Hotes, S. & Ekschmitt, K. Influence of grain size on species–habitat models. Ecol. Modell. 222, 3403–3412 (2011).Article 

    Google Scholar 
    Zuur, A. F., Ieno, E. N. & Elphick, C. S. A protocol for data exploration to avoid common statistical problems. Methods Ecol. Evol. 1, 3–14 (2010).Article 

    Google Scholar 
    Passadore, C., Möller, L. M., Diaz-Aguirre, F. & Parra, G. J. Modelling dolphin distribution to inform future spatial conservation decisions in a marine protected area. Sci. Rep. 8, 1–14 (2018).Article 
    CAS 

    Google Scholar 
    Parra, G. J., Schick, R. & Corkeron, P. J. Spatial distribution and environmental correlates of Australian snubfin and Indo-Pacific humpback dolphins. Ecography (Cop.) 29, 396–406 (2006).Article 

    Google Scholar 
    Conrad, O. et al. System for Automated Geoscientific Analyses (SAGA) v. 2.1.4. Geosci. Model Dev. 8, 1991–2007 (2015).R Core Team. R version 3.6.1. at https://www.r-project.org/ (2019).RStudio Team. RStudio: Integrated Develpment for R. at http://rstudio.com/ (2019).Dormann, C. F. et al. Collinearity: A review of methods to deal with it and a simulation study evaluating their performance. Ecography (Cop.) 36, 27–46 (2013).Article 

    Google Scholar 
    Tikhonov, G. et al. Joint species distribution modelling with the r-package Hmsc. Methods Ecol. Evol. 11, 442–447 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Gelman, A. & Rubin, D. B. Inference from iterative simulation using multiple sequences. Stat. Sci. 7, 457–472 (1992).Article 
    MATH 

    Google Scholar 
    Pearce, J. & Ferrier, S. Evaluating the predictive performance of habitat models developed using logistic regression. Ecol. Modell. 133, 225–245 (2000).Article 

    Google Scholar 
    Tjur, T. Coefficients of determination in logistic regression models—A new proposal: The coefficient of discrimination. Am. Stat. 63, 366–372 (2009).Article 
    MathSciNet 
    MATH 

    Google Scholar 
    Syme, J. The behavioural ecology of mixed-species groups of delphinids. PhD Thesis, Flinders University, Adelaide, Australia. (Flinders University, 2023).Wang, J. Y. Bottlenose Dolphin, Tursiops aduncus, Indo-Pacific Bottlenose Dolphin. in Encyclopedia of Marine Mammals (eds. Würsig, B., Thewissen, J. G. M. & Kovacs, K. M.) 125–130 (Elsevier, 2018). https://doi.org/10.1016/B978-0-12-804327-1.00073-X.Parra, G. J. & Jefferson, T. A. Humpback Dolphins. in Encyclopedia of Marine Mammals (eds. Würsig, B., Thewissen, J. G. M. & Kovacs, K. M.) 483–489 (Elsevier, 2018). https://doi.org/10.1016/B978-0-12-804327-1.00153-9.Dröge, E., Creel, S., Becker, M. S. & M’soka, J. Spatial and temporal avoidance of risk within a large carnivore guild. Ecol. Evol. 7, 189–199 (2017).Article 
    PubMed 

    Google Scholar 
    Browning, N. E., Cockcroft, V. G. & Worthy, G. A. J. Resource partitioning among South African delphinids. J. Exp. Mar. Bio. Ecol. 457, 15–21 (2014).Article 

    Google Scholar 
    Kiszka, J. J., Méndez-Fernandez, P., Heithaus, M. R. & Ridoux, V. The foraging ecology of coastal bottlenose dolphins based on stable isotope mixing models and behavioural sampling. Mar. Biol. 161, 953–961 (2014).Article 
    CAS 

    Google Scholar 
    Saayman, G. S. & Tayler, C. K. The socioecology of humpback dolphins (Sousa sp.). in Behavior of Marine Animals Current Perspectives in Research Volume 3: Cetaceans (eds. Winn, H. E. & Olla, B. L.) 165–226 (Springer, 1979).Gowans, S. & Whitehead, H. Distribution and habitat partitioning by small odontocetes in the Gully, a submarine canyon on the Scotian Shelf. Can. J. Zool. 73, 1599–1608 (1995).Article 

    Google Scholar 
    Clua, E. Mixed-species feeding aggregation of dolphins, large tunas and seabirds in the Azores. Aquat. Living Resour. 14, 11–18 (2001).Article 

    Google Scholar 
    Quérouil, S. et al. Why do dolphins form mixed-species associations in the azores?. Ethology 114, 1183–1194 (2008).Article 

    Google Scholar 
    Heithaus, M. R. & Dill, L. M. Food availability and tiger shark predation risk influence bottlenose dolphin habitat use. Ecology 83, 480–491 (2002).Article 

    Google Scholar  More

  • in

    Microbiomes of a disease-resistant genotype of Acropora cervicornis are resistant to acute, but not chronic, nutrient enrichment

    Acropora Biological Review Team. Atlantic Acropora Status Review: Report to National Marine Fisheries Service (Acropora Biological Review Team, 2005).
    Google Scholar 
    Gardner, T. A., Côté, I. M., Gill, J. A., Grant, A. & Watkinson, A. R. Long-term region-wide declines in Caribbean Corals. Science 301, 958–960 (2003).ADS 
    CAS 
    PubMed 

    Google Scholar 
    Jackson, E. J., Donovan, M., Cramer, K. & Lam, V. Status and Trends of Caribbean Coral Reefs: 1970–2012 306 (International Union for the Conservation of Nature, 2012).
    Google Scholar 
    Schopmeyer, S. A. et al. Regional restoration benchmarks for Acropora cervicornis. Coral Reefs 36, 1047–1057 (2017).ADS 

    Google Scholar 
    Lirman, D. et al. Propagation of the threatened staghorn coral Acropora cervicornis: Methods to minimize the impacts of fragment collection and maximize production. Coral Reefs 29, 729–735 (2010).ADS 

    Google Scholar 
    Mercado-Molina, A. E., Ruiz-Diaz, C. P. & Sabat, A. M. Demographics and dynamics of two restored populations of the threatened reef-building coral Acropora cervicornis. J. Nat. Conserv. 24, 17–23 (2015).
    Google Scholar 
    Young, C., Schopmeyer, S. & Lirman, D. A review of reef restoration and coral propagation using the threatened genus Acropora in the Caribbean and Western Atlantic. Bull. Mar. Sci. 88, 1075–1098 (2012).
    Google Scholar 
    Carne, L., Kaufman, L. & Scavo, K. Measuring success for Caribbean acroporid restoration: key results from ten years of work in southern Belize. In Proc. 13th International Coral Reef Symposium, Honolulu (Abstract No. 27909) (2016).Ware, M. et al. Survivorship and growth in staghorn coral (Acropora cervicornis) outplanting projects in the Florida Keys National Marine Sanctuary. PLoS ONE 15, e0231817 (2020).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Shaver, E. C. et al. A roadmap to integrating resilience into the practice of coral reef restoration. Glob. Change Biol. 28, 4751–4764 (2022).CAS 

    Google Scholar 
    DeFilippo, L. B. et al. Assessing the potential for demographic restoration and assisted evolution to build climate resilience in coral reefs. Ecol. Appl. 32, e2650 (2022).PubMed 
    PubMed Central 

    Google Scholar 
    Lapointe, B. E., Brewton, R. A., Herren, L. W., Porter, J. W. & Hu, C. Nitrogen enrichment, altered stoichiometry, and coral reef decline at Looe Key, Florida Keys, USA: A 3-decade study. Mar. Biol. 166, 108 (2019).
    Google Scholar 
    Montenero, K. A. Florida Keys Integrated Ecosystem Assessment Ecosystem Status Report. https://doi.org/10.25923/F7CE-ST38.Palacio-Castro, A. M., Dennison, C. E., Rosales, S. M. & Baker, A. C. Variation in susceptibility among three Caribbean coral species and their algal symbionts indicates the threatened staghorn coral, Acropora cervicornis, is particularly susceptible to elevated nutrients and heat stress. Coral Reefs 40, 1601–1613 (2021).
    Google Scholar 
    Vega Thurber, R. L. et al. Chronic nutrient enrichment increases prevalence and severity of coral disease and bleaching. Glob. Change Biol. 20, 544–554 (2014).ADS 

    Google Scholar 
    Zaneveld, J. R. et al. Overfishing and nutrient pollution interact with temperature to disrupt coral reefs down to microbial scales. Nat. Commun. 7, 11833 (2016).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bruno, J. F. et al. Thermal stress and coral cover as drivers of coral disease outbreaks. PLoS Biol. 5, e124 (2007).PubMed 
    PubMed Central 

    Google Scholar 
    Wiedenmann, J. et al. Nutrient enrichment can increase the susceptibility of reef corals to bleaching. Nat. Clim. Change 3, 160–164 (2012).ADS 

    Google Scholar 
    Rädecker, N., Pogoreutz, C., Voolstra, C. R., Wiedenmann, J. & Wild, C. Nitrogen cycling in corals: The key to understanding holobiont functioning? Trends Microbiol. 23, 490–497 (2015).PubMed 

    Google Scholar 
    Shantz, A. A. & Burkepile, D. E. Context-dependent effects of nutrient loading on the coral–algal mutualism. Ecology 95, 1995–2005 (2014).PubMed 

    Google Scholar 
    Burkepile, D. E. et al. Nitrogen identity drives differential impacts of nutrients on coral bleaching and mortality. Ecosystems 23, 798–811 (2020).CAS 

    Google Scholar 
    Fabricius, K. E. Effects of terrestrial runoff on the ecology of corals and coral reefs: Review and synthesis. Mar. Pollut. Bull. 50, 125–146 (2005).CAS 
    PubMed 

    Google Scholar 
    Ferrier-Pagès, C., Gattuso, J.-P., Dallot, S. & Jaubert, J. Effect of nutrient enrichment on growth and photosynthesis of the zooxanthellate coral Stylophora pistillata. Coral Reefs 19, 103–113 (2000).
    Google Scholar 
    Bourne, D. G., Morrow, K. M. & Webster, N. S. Insights into the coral microbiome: Underpinning the health and resilience of reef ecosystems. Annu. Rev. Microbiol. 70, 317–340 (2016).CAS 
    PubMed 

    Google Scholar 
    Krediet, C. J., Ritchie, K. B., Paul, V. J. & Teplitski, M. Coral-associated micro-organisms and their roles in promoting coral health and thwarting diseases. Proc. R. Soc. B Biol. Sci. 280, 20122328 (2013).
    Google Scholar 
    Mao-Jones, J., Ritchie, K. B., Jones, L. E. & Ellner, S. P. How microbial community composition regulates coral disease development. PLoS Biol. 8, e1000345 (2010).PubMed 
    PubMed Central 

    Google Scholar 
    Zilber-Rosenberg, I. & Rosenberg, E. Role of microorganisms in the evolution of animals and plants: The hologenome theory of evolution. FEMS Microbiol. Rev. 32, 723–735 (2008).CAS 
    PubMed 

    Google Scholar 
    West, A. G. et al. The microbiome in threatened species conservation. Biol. Conserv. 229, 85–98 (2019).
    Google Scholar 
    Ritchie, K. Regulation of microbial populations by coral surface mucus and mucus-associated bacteria. Mar. Ecol. Prog. Ser. 322, 1–14 (2006).ADS 
    CAS 

    Google Scholar 
    Rohwer, F., Seguritan, V., Azam, F. & Knowlton, N. Diversity and distribution of coral-associated bacteria. Mar. Ecol. Prog. Ser. 243, 1–10 (2002).ADS 

    Google Scholar 
    Klinges, G., Maher, R. L., Thurber, R. L. V. & Muller, E. M. Parasitic ‘Candidatus aquarickettsia rohweri’ is a marker of disease susceptibility in Acropora cervicornis but is lost during thermal stress. Environ. Microbiol. 22, 5341–5355 (2020).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Williams, S. D. et al. Geographically driven differences in microbiomes of Acropora cervicornis originating from different regions of Florida’s Coral Reef. PeerJ 10, e13574 (2022).PubMed 
    PubMed Central 

    Google Scholar 
    Klinges, J. G., Patel, S. H., Duke, W. C., Muller, E. M. & Vega Thurber, R. L. Phosphate enrichment induces increased dominance of the parasite Aquarickettsia in the coral Acropora cervicornis. FEMS Microbiol. Ecol. 98, 013 (2022).
    Google Scholar 
    Rosales, S. M. et al. Microbiome differences in disease-resistant vs susceptible Acropora corals subjected to disease challenge assays. Sci. Rep. 9, 18279 (2019).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Gignoux-Wolfsohn, S., Precht, W., Peters, E., Gintert, B. & Kaufman, L. Ecology, histopathology, and microbial ecology of a white-band disease outbreak in the threatened staghorn coral Acropora cervicornis. Dis. Aquat. Org. 137, 217–237 (2020).
    Google Scholar 
    Miller, N., Maneval, P., Manfrino, C., Frazer, T. K. & Meyer, J. L. Spatial distribution of microbial communities among colonies and genotypes in nursery-reared Acropora cervicornis. PeerJ 8, e9635 (2020).PubMed 
    PubMed Central 

    Google Scholar 
    Aguirre, E. G., Million, W. C., Bartels, E., Krediet, C. J. & Kenkel, C. D. Host-specific epibiomes of distinct Acropora cervicornis genotypes persist after field transplantation. Coral Reefs. https://doi.org/10.1007/s00338-022-02218-x (2022).Article 

    Google Scholar 
    Shaver, E. C. et al. Effects of predation and nutrient enrichment on the success and microbiome of a foundational coral. Ecology 98, 830–839 (2017).PubMed 

    Google Scholar 
    Muller, E. M., Bartels, E. & Baums, I. B. Bleaching causes loss of disease resistance within the threatened coral species Acropora cervicornis. eLife 7, e35066 (2018).PubMed 
    PubMed Central 

    Google Scholar 
    Miller, M. W. et al. Genotypic variation in disease susceptibility among cultured stocks of Elkhorn and Staghorn corals. PeerJ 7, e6751 (2019).PubMed 
    PubMed Central 

    Google Scholar 
    Sunagawa, S., Woodley, C. M. & Medina, M. Threatened corals provide underexplored microbial habitats. PLoS ONE 5, e9554 (2010).ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Pantos, O. et al. The bacterial ecology of a plague-like disease affecting the Caribbean coral Montastrea annularis. Environ. Microbiol. 5, 370–382 (2003).CAS 
    PubMed 

    Google Scholar 
    Sheu, S.-Y., Liu, L.-P., Tang, S.-L. & Chen, W.-M. Thalassotalea euphylliae sp. nov., isolated from the torch coral Euphyllia glabrescens. Int. J. Syst. Evol. Microbiol. 66, 5039–5045 (2016).CAS 
    PubMed 

    Google Scholar 
    Nakagawa, T., Iino, T., Suzuki, K.-I. & Harayama, S. Ferrimonas futtsuensis sp. nov. and Ferrimonas kyonanensis sp. nov., selenate-reducing bacteria belonging to the Gammaproteobacteria isolated from Tokyo Bay. Int. J. Syst. Evol. Microbiol. 56, 2639–2645 (2006).CAS 
    PubMed 

    Google Scholar 
    Maher, R. L. et al. Coral microbiomes demonstrate flexibility and resilience through a reduction in community diversity following a thermal stress event. Front. Ecol. Evol. 8, 1 (2020).ADS 

    Google Scholar 
    Bourne, D., Iida, Y., Uthicke, S. & Smith-Keune, C. Changes in coral-associated microbial communities during a bleaching event. ISME J. 2, 350–363 (2008).CAS 
    PubMed 

    Google Scholar 
    Ziegler, M. et al. Coral bacterial community structure responds to environmental change in a host-specific manner. Nat. Commun. 10, 3092 (2019).ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    McDevitt-Irwin, J. M. et al. Responses of coral-associated bacterial communities to local and global stressors. Front. Mar. Sci. 4, 262 (2017).
    Google Scholar 
    Klinges, J. G. et al. Phylogenetic, genomic, and biogeographic characterization of a novel and ubiquitous marine invertebrate-associated Rickettsiales parasite, Candidatus aquarickettsia rohweri, gen. nov., sp. nov. ISME J. 13, 2938–2953 (2019).PubMed 
    PubMed Central 

    Google Scholar 
    Muscatine, L., Falkowski, P. G., Dubinsky, Z., Cook, P. A. & McCloskey, L. R. The effect of external nutrient resources on the population dynamics of zooxanthellae in a reef coral. Proc. R. Soc. Lond. B 236, 311–324 (1989).ADS 

    Google Scholar 
    Waite, D. W. et al. Comparative genomic analysis of the class Epsilonproteobacteria and proposed reclassification to Epsilonbacteraeota (phyl. Nov.). Front. Microbiol. 8, 682 (2017).PubMed 
    PubMed Central 

    Google Scholar 
    Waite, D. W. et al. Addendum: Comparative genomic analysis of the class Epsilonproteobacteria and proposed reclassification to Epsilonbacteraeota (phyl. Nov.). Front. Microbiol. 9, 772 (2018).PubMed 
    PubMed Central 

    Google Scholar 
    Rosales, S. M. et al. Bacterial metabolic potential and micro-eukaryotes enriched in stony coral tissue loss disease lesions. Front. Mar. Sci. 8, 776859 (2022).
    Google Scholar 
    Ricci, F. et al. Beneath the surface: Community assembly and functions of the coral skeleton microbiome. Microbiome 7, 159 (2019).PubMed 
    PubMed Central 

    Google Scholar 
    Yang, S.-H. et al. Metagenomic, phylogenetic, and functional characterization of predominant endolithic green sulfur bacteria in the coral Isopora palifera. Microbiome 7, 3 (2019).PubMed 
    PubMed Central 

    Google Scholar 
    Cai, L. et al. Metagenomic analysis reveals a green sulfur bacterium as a potential coral symbiont. Sci. Rep. 7, 9320 (2017).ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Allgeier, J. E., Burkepile, D. E. & Layman, C. A. Animal pee in the sea: Consumer-mediated nutrient dynamics in the world’s changing oceans. Glob. Change Biol. 23, 2166–2178 (2017).ADS 

    Google Scholar 
    Hughes, D. J. et al. Coral reef survival under accelerating ocean deoxygenation. Nat. Clim. Change 10, 296–307 (2020).ADS 
    CAS 

    Google Scholar 
    Miura, N. et al. Ruegeria sp. strains isolated from the reef-building coral Galaxea fascicularis inhibit growth of the temperature-dependent pathogen Vibrio coralliilyticus. Mar. Biotechnol. 21, 1–8 (2019).CAS 

    Google Scholar 
    Bruno, J. F., Petes, L. E., Harvell, C. D. & Hettinger, A. Nutrient enrichment can increase the severity of coral diseases. Ecol. Lett. 6, 1056–1061 (2003).
    Google Scholar 
    Ezzat, L. et al. Thermal stress interacts with surgeonfish feces to increase coral susceptibility to dysbiosis and reduce tissue regeneration. Front. Microbiol. 12, 620458 (2021).PubMed 
    PubMed Central 

    Google Scholar 
    Gajigan, A. P., Diaz, L. A. & Conaco, C. Resilience of the prokaryotic microbial community of Acropora digitifera to elevated temperature. Microbiol. Open 6, e00478 (2017).
    Google Scholar 
    MacKnight, N. J. et al. Microbial dysbiosis reflects disease resistance in diverse coral species. Commun. Biol. 4, 679 (2021).PubMed 
    PubMed Central 

    Google Scholar 
    Palacio-Castro, A. M., Rosales, S. M., Dennison, C. E. & Baker, A. C. Microbiome signatures in Acropora cervicornis are associated with genotypic resistance to elevated nutrients and heat stress. Coral Reefs 41, 1389–1403 (2022).
    Google Scholar 
    Vollmer, S. V. & Kline, D. I. Natural disease resistance in threatened staghorn corals. PLoS ONE 3, e3718 (2008).ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hughes, T. P. et al. Spatial and temporal patterns of mass bleaching of corals in the Anthropocene. Science 359, 80–83 (2018).ADS 
    CAS 
    PubMed 

    Google Scholar 
    Parkinson, J. E. et al. Extensive transcriptional variation poses a challenge to thermal stress biomarker development for endangered corals. Mol. Ecol. 27, 1103–1119 (2018).CAS 
    PubMed 

    Google Scholar 
    Siebeck, U. E., Logan, D. & Marshall, N. J. CoralWatch—A flexible coral bleaching monitoring tool for you and your group. In Proc. 11th Int. Coral Reef Symp. Ft Lauderdale, Florida, 7–11 July, Vol. 1392, 5 (2008).Parada, A. E., Needham, D. M. & Fuhrman, J. A. Every base matters: Assessing small subunit rRNA primers for marine microbiomes with mock communities, time series and global field samples. Environ. Microbiol. 18, 1403–1414 (2016).CAS 
    PubMed 

    Google Scholar 
    Apprill, A., McNally, S., Parsons, R. & Weber, L. Minor revision to V4 region SSU rRNA 806R gene primer greatly increases detection of SAR11 bacterioplankton. Aquat. Microb. Ecol. 75, 129–137 (2015).
    Google Scholar 
    Messyasz, A., Maher, R. L., Meiling, S. S. & Thurber, R. V. Nutrient enrichment predominantly affects low diversity microbiomes in a marine trophic symbiosis between algal farming fish and corals. Microorganisms 9, 1873 (2021).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet. J. 17, 10–12 (2011).
    Google Scholar 
    Callahan, B. J. et al. DADA2: High-resolution sample inference from Illumina amplicon data. Nat. Methods 13, 581–583 (2016).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    McMurdie, P. J. & Holmes, S. phyloseq: An R package for reproducible interactive analysis and graphics of microbiome census data. PLoS ONE 8, e61217 (2013).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Magurran, A. E. Ecological Diversity and Its Measurement (Princeton University Press, 1988).
    Google Scholar 
    Lahti, L. & Shetty, S. Microbiome R Package.Gloor, G. B., Macklaim, J. M., Pawlowsky-Glahn, V. & Egozcue, J. J. Microbiome datasets are compositional: And this is not optional. Front. Microbiol. 8, 2224 (2017).PubMed 
    PubMed Central 

    Google Scholar 
    Anderson, M. J. A new method for non-parametric multivariate analysis of variance. Austral Ecol. 26, 32–46 (2001).
    Google Scholar 
    Oksanen, J. et al. vegan: Community Ecology Package (2019).Martinez Arbizu, P. pairwiseAdonis: Pairwise multilevel comparison using adonis. R Package Version 0.0.1 (2017).Anderson, M. J. Distance-based tests for homogeneity of multivariate dispersions. Biometrics 62, 245–253 (2006).MathSciNet 
    PubMed 
    MATH 

    Google Scholar 
    Kaul, A., Mandal, S., Davidov, O. & Peddada, S. D. Analysis of microbiome data in the presence of excess zeros. Front. Microbiol. 8, 2114 (2017).PubMed 
    PubMed Central 

    Google Scholar  More

  • in

    Sand fly population dynamics in areas of American cutaneous leishmaniasis, Municipality of Paraty, Rio de Janeiro, Brazil

    Owing to drastic changes in the environment caused by human interference, wild mammals that are reservoirs of Leishmania have invaded residential areas where species of sand flies with eclectic feeding habits are found, and established a transmission cycle that eventually reaches humans23,24,25. In the study area, it was observed that the largest frequency of specimens over the years was captured in the residential environment, which are represented by residential and peridomicile areas. The lowest frequency was captured in the borders of the forest.The municipality of Paraty, located on the southern coast in the state of Rio de Janeiro, where the study was conducted, has many preserved areas of the Atlantic Forest and its climate is wet with no dry season13, which was confirmed during the three years of the present study, where the relative air humidity stayed high every month. The highest average rainfalls occur in summer and fall (autumn). The average temperature during the hottest months of the year was between approximately 25 °C and 26 °C, with a maximum of 31 °C, and in the coldest months, the temperature averaged between 20 and 21 °C, with a minimum of 16 °C, exhibiting an ideal environment for the activity of sand flies throughout the year.Barretto26 noted that atmospheric conditions, such as relative humidity, rainfall, and temperature directly influence the activity of these sand fly species. Migonemyia migonei and Ny. whitmani had lower activity at temperatures below 15 °C, Pi. fischeri below 10 °C, and Ny. intermedia at temperatures below 9.5 °C. The author also reported that heavy rains prevent sand flies from leaving their shelters; however, this can increase their density within residences, especially for species located next to residential areas. Light rain will not impede their activity, but in these conditions, they are not as frequently observed as they usually are. However, during rain periods, especially in the hot and humid summer period, the density of sand flies increases considerably.In the present study, four key vector species of Leishmania braziliensis Vianna, 1911, the etiologic agent of tegumentary leishmaniasis, were captured throughout the year. The most frequent was Ny. intermedia, followed by Pi. fischeri, Mg. migonei, and Ny. whitmani. Carvalho et al.27, in the State of Pernambuco, northeast region of Brazil, reported having found Mg. migonei infected with Leishmania infantum Nicolle, 1908, the etiologic agent of visceral leishmaniasis.According to Forattini28, there are sand fly species that are essentially resistant to climate changes throughout the seasons. Several are found, albeit in lower densities, during the cooler, dry months, while others disappear during this period. However, other factors also influence the incidence of sand flies in the same location, even under the same temperature and humidity conditions. Thus, to study the seasonality of sand fly species, it is important to perform systematized captures, for a period exceeding two years, to minimize the effects of these additional factors, for example, atypical years with a longer period of drought or humidity, more or less high temperatures, months with higher than expected rainfall or control measures applied by the municipality.In studies carried out in the Northeast region of Brazil, in a study carried out in the municipality of Codó, in the State of Maranhão, an inversely proportional correlation of the captured sandflies was observed in relation to relative air humidity, a direct correlation in relation to temperature and precipitation, a correlation directly proportional29. In the municipality of Sobral, State of Ceará, in the first year of the study, observed a negative correlation with temperature and a high positive correlation with humidity and precipitation, however, in the following year, there was no correlation between the density of captured sandflies and climatic variables30. The same occurred in this study, in the municipality of Paraty, in relation to relative air humidity and precipitation, but in relation to temperature, a strong positive correlation was obtained.In the studied area Ny. intermedia occurred in greater numbers in every month of the year, except in June and July, when it was less frequent than Pi. fischeri. The same pattern was observed for these two species, i.e., a gradual increase in abundance beginning in August, peak abundance in summer (January), followed by a decrease until winter (July). Brito et al.31, when researching the northern coast of the state of São Paulo, municipality of São Sebastião, noted the opposite, that Ny. intermedia had the highest abundance peaks during the driest and coldest period of the year, i.e., from May to August. However, the authors also emphasized the presence of this species throughout the year, mainly in the residential environment, and they stressed the importance of seasonal analyses for periods longer than a year.In the São Francisco River region, in the state of Minas Gerais, on the banks of the Rio Velhas, Saraiva et al.32, in a study over a two-year period, observed a different pattern. In the first year of study, after the rainy season from February to May, with high humidity and high temperature, Ny. intermedia was captured in greater numbers than during other months of the year. In the second year, peaks occurred in October, March, and June, with the highest peak in March, when there was elevated rainfall, high humidity, and high temperatures.In the state of Rio de Janeiro, in Serra dos Órgãos National Park, Aguiar and Soucasaux33 analyzed the monthly frequency in human bait and observed that Ny. fischeri was captured in every month except November. In the hot and humid period, from December to February, there was a gradual increase in the average abundances of this species, and then a slight decrease began in March and continued into April. During the cold and dry period of May and June, abundances started to increase, then decreased in July, and peaked in August. During August, Pi. fischeri was the dominant species of wildlife, and in September, abundances began to decline again.Mayo et al.34, studying the southeastern region of the state of São Paulo, observed that there was a seasonal trend in the abundance for species Mg. migonei, Ny. whitmani, Ny. intermedia, and Pi. fischeri, with abundance peaks recorded during the cooler, drier season (April to September) and low abundances during the warmer, wetter season (October to March). The authors revealed that the occurrence of intense fires in the study area in October, which caused severe environmental change, possibly interfered with the population dynamics of the species. In the present study, the opposite trend of seasonality was shown for the four key species, Ny. intermedia, Pi. fischeri, Mg. migonei, and Ny. whitmani, then what was observed by the above authors, the highest abundances occurred during the hottest period, increasing gradually until a maximum peak in January, and lowest abundances were seen during the coldest period, in July for the first three species, and in June for Ny. whitmani.In the neighboring municipality of this study in Angra dos Reis, in the Ilha Grande, Carvalho et al.35 reinforced the epidemiological importance of Ny. intermedia in the State of Rio de Janeiro and highlighted the role of Mg. migonei in the transmission of cutaneous leishmaniasis with its high rate of infection natural by Leishmania. Still in the same region, along the southern coast of the State of Rio de Janeiro, Aguiar et al.8 conducted systematic catches for two years, with the aim being to analyze the monthly frequency of sand flies in residential and forest environments. The authors discovered results like what occurred in this study in Paraty, that the four most important species caught, Ny. intermedia, Pi. fischeri, Mg. migonei, and Ny. whitmani, had higher average numbers during the hot and humid period of the year, i.e., between October and January, with a maximum peak in December for Ny. intermedia and Pi. fischeri, and January for Mg. migonei. The prevalence of Ny. intermedia was evident in every month, both inside the residence and around the residential area. In the colder and drier season, from May to August, there was a balance with Pi. fischeri, but from August, inside the residence, and from September, around the residence, the frequency increased until it reached its peak in December. There was a gradual increase in the frequency of this species in the warmer and wetter period (between October and January), with average temperatures ranging from 26 to 29 °C and relative air humidity between 84 and 87%.Condino et al.36, when studying the southwestern region of the state of São Paulo, observed that Ny. intermedia and Ny. whitmani had the highest frequencies during the months of May, September, and December with temperatures ranging from 21 to 25.7 °C and rainfall between 66.7 and 195.1 mm. In June, the lowest frequency of sand flies was observed, which then increased until a maximum peak in September. Temperature data and rainfall index were not correlated with the density of specimens, especially as the study was carried out over only one year. In this study, the opposite was observed for Ny. intermedia and Ny. whitmani in the month of May, one of the months with the lowest density.In the city of Petrópolis, state of Rio de Janeiro, Souza et al.24 observed a prevalence of Ny. intermedia and Ny. whitmani, with the latter species prevailing around the residence. Migonemyia migonei and Pi. fischeri were also present but to a lesser extent. In the forest, Ny. whitmani was more abundant, followed by Pi. fischeri, while Ny. intermedia was found at lower abundances. However, Ny. intermedia and Pi. fischeri were present during every month of the year. The authors also found a significant correlation between the number of sand flies and environmental changes such as temperature, relative humidity, and rainfall. The same was observed, in this study, in the forest with Ny. intermedia, however, in this environment the number of Pi. fischeri specimens was higher than that of Ny. whitmani.In the north of Espírito Santo, Virgens et al.37 observed that Ny. intermedia was present in almost every month of the study period, with peaks in the warmer and wetter months. The authors highlighted that the low numbers of this species were recorded during and after high rainfall periods, suggesting that heavy rain is unfavorable for the development of immature forms, as breeding sites in altered habitats suffered a greater impact because of extreme weather conditions.In a study carried out by Guimarães et al.38 to observe the competence of Mg. Migonei to Leishmania infantum, concluded that this species is highly susceptible to the development of this parasite and that in addition to its anthropophilia and abundance in areas with an active focus of visceral leishmaniasis, it can act as a vector of this disease in Latin America.In the studied area, Ny. intermedia, one of the main vectors of the etiological agent of tegumentary leishmaniasis in the region2, was present in significant numbers in the home environment throughout all months of the year. The species Pi. fischeri was present over the months in expressive numbers in all types and locations of capture, that is, both in the environment altered by human activity and in the natural environment where leishmaniasis occurs in its natural enzootic cycle. Migonemyia migonei, present throughout the year in the peridomestic environment, showed its association with the dog, where it was prevalent throughout the year in the kennel, being an important vector of the etiological agent of tegumentary leishmaniasis, as well as being suspected in areas of visceral leishmaniasis transmission, where the main vector of this disease is not found. And Ny. whitmani present in the peridomicile, mainly in the hottest months of the year, in addition to the forest and forest margins, it was observed that in this study region the species is emerging through a selective process of adaptation in environments that were negatively affected by the increase of human activity. Thus, despite observing a period of greater frequency of sand flies in the hottest months of the year, a period with high rainfall, the high relative humidity is observed throughout the year, as well as the presence of species of epidemiological importance Ny. intermedia, Pi. fischeri, Mg. migonei and Ny. whitmani, who are involved in the propagation of the etiological agent of tegumentary leishmaniasis to humans and animals, causing greater contact between the region’s inhabitants with these dipterans and thus, a greater risk of contracting the disease. More

  • in

    Integrating multiple plant functional traits to predict ecosystem productivity

    Beer, C. et al. Terrestrial gross carbon dioxide uptake: global distribution and covariation with climate. Science 329, 834–838 (2010).Article 
    CAS 
    PubMed 

    Google Scholar 
    Badgley, G., Field, C. B. & Berry, J. A. Canopy near-infrared reflectance and terrestrial photosynthesis. Sci. Adv.3, e1602244 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Chapin, F. S. III Effects of plant traits on ecosystem and regional processes: a conceptual framework for predicting the consequences of global change. Ann. Bot. 91, 455–463 (2003).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Chu, C. et al. Does climate directly influence NPP globally? Global Chan. Biol. 22, 12–24 (2016).Article 

    Google Scholar 
    Yao, Y. et al. Spatiotemporal pattern of gross primary productivity and its covariation with climate in China over the last thirty years. Global Chan. Biol. 24, 184–196 (2018).Article 

    Google Scholar 
    Fang, J., Lutz, J. A., Wang, L., Shugart, H. H. & Yan, X. Using climate-driven leaf phenology and growth to improve predictions of gross primary productivity in North American forests. Global Chan. Biol. 26, 6974–6988 (2020).Article 

    Google Scholar 
    Fernández-Martínez, M. et al. The role of climate, foliar stoichiometry and plant diversity on ecosystem carbon balance. Global Chan. Biol. 26, 7067–7078 (2020).Article 

    Google Scholar 
    Migliavacca, M. et al. The three major axes of terrestrial ecosystem function. Nature 598, 468–472 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Reichstein, M., Bahn, M., Mahecha, M. D., Kattge, J. & Baldocchi, D. D. Linking plant and ecosystem functional biogeography. Proc. Natl. Acad. Sci. 111, 13697–13702 (2014).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Funk, J. L. et al. Revisiting the Holy Grail: using plant functional traits to understand ecological processes. Biol. Rev. 92, 1156–1173 (2017).Article 
    PubMed 

    Google Scholar 
    Lavorel, S. & Garnier, É. Predicting changes in community composition and ecosystem functioning from plant traits: revisiting the Holy Grail. Fun. Ecol. 16, 545–556 (2002).Article 

    Google Scholar 
    Enquist, B. J. et al. in Advances in Ecological Research 52 (eds Samraat P, Guy W, & Anthony I. D) 249–318 (Academic Press, 2015).Garnier, E. et al. Plant functional markers capture ecosystem properties during secondary succession. Ecology 85, 2630–2637 (2004).Article 

    Google Scholar 
    Enquist, B. J. et al. Assessing trait-based scaling theory in tropical forests spanning a broad temperature gradient. Global Ecol. Biogeogr. 26, 1357–1373 (2017).Article 

    Google Scholar 
    Fyllas, N. M. et al. Solar radiation and functional traits explain the decline of forest primary productivity along a tropical elevation gradient. Ecol. Lett. 20, 730–740 (2017).Article 
    PubMed 

    Google Scholar 
    Van der Plas, F. et al. Plant traits alone are poor predictors of ecosystem properties and long-term ecosystem functioning. Nat. Ecol. Evol. 4, 1602–1611 (2020).Article 
    PubMed 

    Google Scholar 
    Ali, A., Yan, E.-R., Chang, S. X., Cheng, J.-Y. & Liu, X.-Y. Community-weighted mean of leaf traits and divergence of wood traits predict aboveground biomass in secondary subtropical forests. Sci. Total Environ. 574, 654–662 (2017).Article 
    CAS 
    PubMed 

    Google Scholar 
    Yang, J., Cao, M. & Swenson, N. G. Why Functional Traits Do Not Predict Tree Demographic Rates. Trend Ecol. Evol. 33, 326–336 (2018).Article 

    Google Scholar 
    Šímová, I. et al. The relationship of woody plant size and leaf nutrient content to large-scale productivity for forests across the Americas. J. Ecol. 107, 2278–2290 (2019).Article 

    Google Scholar 
    Li, Y. et al. Leaf size of woody dicots predicts ecosystem primary productivity. Ecol. Lett. 23, 1003–1013 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    He, N. et al. Ecosystem Traits Linking Functional Traits to Macroecology. Trend. Ecol. Evol. 34, 200–210 (2019).Article 

    Google Scholar 
    Rubio, V. E., Zambrano, J., Iida, Y., Umaña, M. N. & Swenson, N. G. Improving predictions of tropical tree survival and growth by incorporating measurements of whole leaf allocation. J. Ecol. 109, 1331–1343 (2021).Article 

    Google Scholar 
    Drake, J. E. et al. Increases in the flux of carbon belowground stimulate nitrogen uptake and sustain the long-term enhancement of forest productivity under elevated CO2. Ecol. Lett. 14, 349–357 (2011).Article 
    PubMed 

    Google Scholar 
    Hilty, J., Muller, B., Pantin, F. & Leuzinger, S. Plant growth: the What, the How, and the Why. New Phytol. 232, 25–41 (2021).Article 
    PubMed 

    Google Scholar 
    Xia, J. et al. Joint control of terrestrial gross primary productivity by plant phenology and physiology. Proc. Natl. Acad. Sci. 112, 2788–2793 (2015).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Suding, K. N. et al. Scaling environmental change through the community-level: A trait-based response-and-effect framework for plants. Global Chan. Biol. 14, 1125–1140 (2008).Article 

    Google Scholar 
    Liu, C., Li, Y., Yan, P. & He, N. How to Improve the Predictions of Plant Functional Traits on Ecosystem Functioning? Front. Plant Sci. 12, 622260 (2021).Reich, P. B., Walters, M. B. & Ellsworth, D. S. From tropics to tundra: global convergence in plant functioning. Proc. Natl. Acad. of Sci. 94, 13730–13734 (1997).Article 
    CAS 

    Google Scholar 
    Reich, P. B. The world-wide ‘fast–slow’ plant economics spectrum: a traits manifesto. J. Ecol. 102, 275–301 (2014).Article 

    Google Scholar 
    Monteith, J. L. Climate and the efficiency of crop production in Britain. Philosophical Transactions of the Royal Society of London. B. Biol. Sci. 281, 277–294 (1977).
    Google Scholar 
    Garnier, E. Resource capture, biomass allocation and growth in herbaceous plants. Trend. Ecol. Evol. 6, 126–131 (1991).Article 
    CAS 

    Google Scholar 
    Farnsworth, K. D., Albantakis, L. & Caruso, T. Unifying concepts of biological function from molecules to ecosystems. Oikos 126, 1367–1376 (2017).Article 

    Google Scholar 
    Zhang, R. et al. Biodiversity alleviates the decrease of grassland multifunctionality under grazing disturbance: A global meta-analysis. Global Ecol. Biogeogr. 31, 155–167 (2022).Article 

    Google Scholar 
    Jing, X. et al. The links between ecosystem multifunctionality and above-and belowground biodiversity are mediated by climate. Nat. Commun. 6, 1–8 (2015).Article 

    Google Scholar 
    Peters, M. K. et al. Climate–land-use interactions shape tropical mountain biodiversity and ecosystem functions. Nature 568, 88–92 (2019).Article 
    CAS 
    PubMed 

    Google Scholar 
    Hu, W. et al. Aridity-driven shift in biodiversity–soil multifunctionality relationships. Nat. Commun. 12, 1–15 (2021).Article 

    Google Scholar 
    Jing, X. et al. The influence of aboveground and belowground species composition on spatial turnover in nutrient pools in alpine grasslands. Global Ecol. Biogeogr. 31, 486–500 (2022).Article 

    Google Scholar 
    Jing, X. et al. Above-and belowground complementarity rather than selection drives tree diversity-productivity relationships in European forests. Funct Ecol. 35, 1756–1767 (2021).He, N. et al. Predicting ecosystem productivity based on plant community traits. Trend. Plant Sci. 28, 43–53 (2023).Maynard, D. S. et al. Global relationships in tree functional traits. Nat. Commun. 13, 1–12 (2022).Article 

    Google Scholar 
    Michaletz, S. T., Kerkhoff, A. J. & Enquist, B. J. Drivers of terrestrial plant production across broad geographical gradients. Global Ecol. Biogeogr. 27, 166–174 (2018).Article 

    Google Scholar 
    Shipley, B. Net assimilation rate, specific leaf area and leaf mass ratio: which is most closely correlated with relative growth rate? A meta-analysis. Funct. Ecol. 20, 565–574 (2006).Article 

    Google Scholar 
    Violle, C. et al. Let the concept of trait be functional! Oikos 116, 882–892 (2007).Article 

    Google Scholar 
    Jucker, T., Bouriaud, O. & Coomes, D. A. Crown plasticity enables trees to optimize canopy packing in mixed-species forests. Funct. Ecol. 29, 1078–1086 (2015).Article 

    Google Scholar 
    McGill, B. J. Matters of Scale. Science 328, 575 (2010).Article 
    CAS 
    PubMed 

    Google Scholar 
    Penuelas, J. et al. Increasing atmospheric CO2 concentrations correlate with declining nutritional status of European forests. Communi. Biol. 3, 1–11 (2020).
    Google Scholar 
    Weemstra, M. et al. Towards a multidimensional root trait framework: a tree root review. New Phytol. 211, 1159–1169 (2016).Article 
    CAS 
    PubMed 

    Google Scholar 
    Oehri, J., Schmid, B., Schaepman-Strub, G. & Niklaus, P. A. Biodiversity promotes primary productivity and growing season lengthening at the landscape scale. Proc. Natl. Acad. Sci. 114, 10160–10165 (2017).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wright, I. J. et al. The worldwide leaf economics spectrum. Nature 428, 821–827 (2004).Article 
    CAS 
    PubMed 

    Google Scholar 
    Diaz, S. et al. The global spectrum of plant form and function. Nature 529, 167–171 (2015).Article 
    PubMed 

    Google Scholar 
    Liu, Y. et al. The optimum temperature of soil microbial respiration: Patterns and controls. Soil Biol. Biochem. 121, 35–42 (2018).Article 
    CAS 

    Google Scholar 
    Zhao, N. et al. Coordinated pattern of multi-element variability in leaves and roots across Chinese forest biomes. Global Ecol. Biogeogr. 25, 359–367 (2016).Article 

    Google Scholar 
    Zhang, J. et al. C: N: P stoichiometry in China’s forests: From organs to ecosystems. Funct. Ecol. 32, 50–60 (2018).Article 

    Google Scholar 
    Karger, D. N. et al. Climatologies at high resolution for the earth’s land surface areas. Sci. Data 4, 1–20 (2017).Article 

    Google Scholar 
    Dirk Nikolaus, K. et al. Climatologies at high resolution for the earth’s land surface areas. EnviDat. https://doi.org/10.16904/envidat.332 (2021).Kerkhoff, A. J., Enquist, B. J., Elser, J. J. & Fagan, W. F. Plant allometry, stoichiometry and the temperature-dependence of primary productivity. Global Ecol. Biogeogr. 14, 585–598 (2005).Article 

    Google Scholar 
    Wright, I. J. et al. Global climatic drivers of leaf size. Science 357, 917–921 (2017).Article 
    CAS 
    PubMed 

    Google Scholar 
    Zhang, Y. et al. A global moderate resolution dataset of gross primary production of vegetation for 2000-2016. Sci. Data 4, 170165 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Jolliffe, I. T. & Cadima, J. Principal component analysis: a review and recent developments. Philos. Trans. Soc. A Math. Phys. Eng. Sci. 374, 20150202 (2016).Article 

    Google Scholar 
    Wieczynski, D. J. et al. Climate shapes and shifts functional biodiversity in forests worldwide. Proc. Natl. Acad. Sci. 116, 587–592 (2019).Article 
    CAS 
    PubMed 

    Google Scholar 
    Lefcheck, J. S. piecewiseSEM: Piecewise structural equation modelling in r for ecology, evolution, and systematics. Method Ecol. Evol. 7, 573–579 (2016).Article 

    Google Scholar 
    Bürkner, P.-C. Advanced bayesian multilevel modeling with the R Package brms. R J. 10, 395–411 (2018).Bürkner, P.-C. brms: An R package for Bayesian multilevel models using Stan. J. Stat. Software 80, 1–28 (2017).Article 

    Google Scholar 
    Vehtari, A., Gelman, A. & Gabry, J. Practical Bayesian model evaluation using leave-one-out cross-validation and WAIC. Stat. Comput. 27, 1413–1432 (2017).Article 

    Google Scholar 
    Vehtari, A. et al. loo: Efficient leave-one-out cross-validation and WAIC for Bayesian models. R package version 2, 1003 (2019).
    Google Scholar 
    Gabry, J. & Mahr, T. bayesplot: Plotting for Bayesian models. R package version 1 (2017).Mac Nally, R. & Walsh, C. J. Hierarchical partitioning public-domain software. Biodivers. Conserv. 13, 659 (2004).Article 

    Google Scholar 
    Murray, K. & Conner, M. M. Methods to quantify variable importance: implications for the analysis of noisy ecological data. Ecology 90, 348–355 (2009).Article 
    PubMed 

    Google Scholar 
    Yan, P., He, N., Yu, K., Xu, L. & Van Meerbeek, K. Integrating multiple functional traits to predict ecosystem productivity. figshare (2023). Dataset. https://doi.org/10.6084/m9.figshare.22081634.v1. More

  • in

    Living in human-modified landscapes narrows the dietary niche of a specialised mammalian scavenger

    Purvis, A., Gittleman, J. L., Cowlishaw, G. & Mace, G. M. Predicting extinction risk in declining species. Proc. R. Soc. B 267, 1947–1952 (2000).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Crooks, K. R. Relative sensitivities of mammalian carnivores to habitat fragmentation. Conserv. Biol. 16, 488–502 (2002).Article 

    Google Scholar 
    Fahrig, L. Non-optimal animal movement in human-altered landscapes. Funct. Ecol. 21, 1003–1015 (2007).Article 

    Google Scholar 
    Fahrig, L. & Rytwinski, T. Effects of roads on animal abundance: An empirical review and synthesis. Ecol. Soc. 14, 21 (2009).Article 

    Google Scholar 
    Lowry, H., Lill, A. & Wong, B. B. M. Behavioural responses of wildlife to urban environments. Biol. Rev. 88, 537–549 (2013).Article 
    PubMed 

    Google Scholar 
    Sévêque, A., Gentle, L. K., López-Bao, J. V., Yarnell, R. W. & Uzal, A. Human disturbance has contrasting effects on niche partitioning within carnivore communities. Biol. Rev. 95, 1689–1705 (2020).Article 
    PubMed 

    Google Scholar 
    Woodroffe, R. & Ginsberg, J. R. Edge effects and the extinction of populations inside protected areas. Science 1979(280), 2126–2128 (1998).Article 
    ADS 

    Google Scholar 
    Dressel, S., Sandström, C. & Ericsson, G. A meta-analysis of studies on attitudes toward bears and wolves across Europe 1976–2012. Conserv. Biol. 29, 565–574 (2015).Article 
    CAS 
    PubMed 

    Google Scholar 
    Owen, D. & Pemberton, D. Tasmanian Devil: A Unique and Threatened Animal (Allen & Unwin, 2005).
    Google Scholar 
    Yirga, G. et al. Adaptability of large carnivores to changing anthropogenic food sources: diet change of spotted hyena (Crocuta crocuta) during Christian fasting period in northern Ethiopia. J. Anim. Ecol. 81, 1052–1055 (2012).Article 
    PubMed 

    Google Scholar 
    Knight, R. L. & Kawashima, J. Y. Responses of raven and red-tailed hawk populations to linear right-of-ways. J. Wildl. Manag. 57, 266–271 (1993).Article 

    Google Scholar 
    Wilmers, C. C., Stahler, D. R., Crabtree, R. L., Smith, D. W. & Getz, W. M. Resource dispersion and consumer dominance: Scavenging at wolf- and hunter-killed carcasses in Greater Yellowstone, USA. Ecol. Lett. 6, 996–1003 (2003).Article 

    Google Scholar 
    Lambertucci, S. A., Speziale, K. L., Rogers, T. E. & Morales, J. M. How do roads affect the habitat use of an assemblage of scavenging raptors?. Biodivers. Conserv. 18, 2063–2074 (2009).Article 

    Google Scholar 
    Šálek, M., Kreisinger, J., Sedláček, F. & Albrecht, T. Do prey densities determine preferences of mammalian predators for habitat edges in an agricultural landscape?. Landsc. Urban Plan. 98, 86–91 (2010).Article 

    Google Scholar 
    Bateman, P. W. & Fleming, P. A. Big city life: Carnivores in urban environments. J. Zool. 287, 1–23 (2012).Article 

    Google Scholar 
    Auman, H. J., Meathrel, C. E. & Richardson, A. Supersize me: Does anthropogenic food change the body condition of silver gulls? A comparison between urbanized and remote, non-urbanized areas. Waterbirds 31, 122–126 (2008).Article 

    Google Scholar 
    Coon, C. A. C., Nichols, B. C., McDonald, Z. & Stoner, D. C. Effects of land-use change and prey abundance on the body condition of an obligate carnivore at the wildland-urban interface. Landsc. Urban Plan. 192, 103648 (2019).Article 

    Google Scholar 
    Beckmann, J. P. & Berger, J. Using black bears to test ideal-free distribution models experimentally. J. Mammal. 84, 594–606 (2003).Article 

    Google Scholar 
    Fedriani, J. M., Fuller, T. K. & Sauvajot, R. M. Does availability of anthropogenic food enhance densities of omnivorous mammals? An example with coyotes in southern California. Ecography 24, 325–331 (2001).Article 

    Google Scholar 
    Prange, S., Gehrt, S. D. & Wiggers, E. P. Influences of anthropogenic resources on raccoon (Procyon lotor) movements and spatial distribution. J. Mammal. 85, 483–490 (2004).Article 

    Google Scholar 
    Tucker, M. A., Santini, L., Carbone, C. & Mueller, T. Mammal population densities at a global scale are higher in human-modified areas. Ecography 44, 1–13 (2021).Article 

    Google Scholar 
    Blanco, G., Lemus, J. A. & García-Montijano, M. When conservation management becomes contraindicated: Impact of food supplementation on health of endangered wildlife. Ecol. Appl. 21, 2469–2477 (2011).Article 
    PubMed 

    Google Scholar 
    Fischer, J. R., Stallknecht, D. E., Luttrell, M. P., Dhondt, A. A. & Converse, K. A. Mycoplasmal conjunctivitis in wild songbirds: The spread of a new contagious disease in a mobile host population. Emerg. Infect. Dis. 3, 69–72 (1997).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Brittingham, M. C. & Temple, S. A. A survey of avian mortality at winter feeders. Wildl. Soc. Bull. 14, 445–450 (1986).
    Google Scholar 
    Hivert, L. G. et al. High blood lead concentrations in captive Tasmanian devils (Sarcophilus harrisii): A threat to the conservation of the species?. Aust. Vet. J. 96, 442–449 (2018).Article 
    CAS 
    PubMed 

    Google Scholar 
    Carrete, M., Donázar, J. A. & Margalida, A. Density-dependent productivity depression in pyrenean bearded vultures: Implications for conservation. Ecol. Appl. 16, 1674–1682 (2006).Article 
    PubMed 

    Google Scholar 
    Bozek, C. K., Prange, S. & Gehrt, S. D. The influence of anthropogenic resources on multi-scale habitat selection by raccoons. Urban Ecosyst. 10, 413–425 (2007).Article 

    Google Scholar 
    Jones, J. D. et al. Supplemental feeding alters migration of a temperate ungulate. Ecol. Appl. 24, 1769–1779 (2014).Article 
    PubMed 

    Google Scholar 
    Šálek, M., Drahníková, L. & Tkadlec, E. Changes in home range sizes and population densities of carnivore species along the natural to urban habitat gradient. Mamm. Rev. 45, 1–14 (2015).Article 

    Google Scholar 
    Newsome, D. & Rodger, K. To feed or not to feed: a contentious issues in wildlife tourism. In Too Close for Comfort: Contentious Issues in Human-Wildlife Encounters (ed. Lunney, D.) 255–270 (Royal Zoological Society of New South Wales, 2008).Chapter 

    Google Scholar 
    Tucker, M. A. et al. Moving in the anthropocene: Global reductions in terrestrial mammalian movements. Science 1979(359), 466–469 (2018).Article 
    ADS 

    Google Scholar 
    Polis, G. A., Anderson, W. B. & Holt, R. D. Toward an integration of landscape and food web ecology: The dynamics of spatially subsidized food webs. Annu. Rev. Ecol. Syst. 28, 289–316 (1997).Article 

    Google Scholar 
    Prange, S. & Gehrt, S. D. Changes in mesopredator-community structure in response to urbanization. Can. J. Zool. 82, 1804–1817 (2004).Article 

    Google Scholar 
    Rodewald, A. D., Kearns, L. J. & Shustack, D. P. Anthropogenic resource subsidies decouple predator–prey relationships. Ecol. Appl. 21, 936–943 (2011).Article 
    PubMed 

    Google Scholar 
    Cortés-Avizanda, A., Jovani, R., Carrete, M. & Donázar, J. A. Resource unpredictability promotes species diversity and coexistence in an avian scavenger guild: A field experiment. Ecology 93, 2570–2579 (2012).Article 
    PubMed 

    Google Scholar 
    Arrondo, E., Cortés-Avizanda, A. & Donázar, J. A. Temporally unpredictable supplementary feeding may benefit endangered scavengers. Ibis 157, 648–651 (2015).Article 

    Google Scholar 
    Smith, J. A., Thomas, A. C., Levi, T., Wang, Y. & Wilmers, C. C. Human activity reduces niche partitioning among three widespread mesocarnivores. Oikos 127, 890–901 (2018).Article 

    Google Scholar 
    de León, L. F. et al. Urbanization erodes niche segregation in Darwin’s finches. Evol. Appl. 12, 1329–1343 (2019).Article 
    PubMed 

    Google Scholar 
    Manlick, P. J. & Pauli, J. N. Human disturbance increases trophic niche overlap in terrestrial carnivore communities. PNAS 117, 26842–26848 (2020).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Blair, R. B. Land use and avian species diversity along an urban gradient. Ecol. Appl. 6, 506–519 (1996).Article 

    Google Scholar 
    Dettori, E. E. et al. Distribution and diet of recovering Eurasian otter (Lutra lutra) along the natural-to-urban habitat gradient (river Segura, SE Spain). Urban Ecosyst. 24, 1221–1230 (2021).Article 

    Google Scholar 
    McKinney, M. L. Urbanization as a major cause of biotic homogenization. Biol. Conserv. 127, 247–260 (2006).Article 

    Google Scholar 
    Guiler, E. R. Temporal and spatial distribution of the Tasmanian Devil, Sarcophilus harrisii (Dasyuridae: Marsupialia). Pap. Proc. R. Soc. Tasman 116, 153–163 (1982).
    Google Scholar 
    Patton, A. H. et al. A transmissible cancer shifts from emergence to endemism in Tasmanian devils. Science (1979) 370, eabb9772 (2020).CAS 

    Google Scholar 
    Cunningham, C. X. et al. Quantifying 25 years of disease-caused declines in Tasmanian devil populations: Host density drives spatial pathogen spread. Ecol. Lett. 24, 958–969 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Rose, R. K., Pemberton, D. A., Mooney, N. J. & Jones, M. E. Sarcophilus harrisii (Dasyuromorphia: Dasyuridae). Mamm. Species 49, 1–17 (2017).Article 

    Google Scholar 
    Guiler, E. R. Observations on the Tasmanian devil, Sarcophilus harrisii (Marsupialia: Dasyuridae) I. Numbers, home range, movements and food in two populations. Aust. J. Zool. 18, 49–62 (1970).Article 

    Google Scholar 
    Jones, M. E. & Barmuta, L. A. Diet overlap and relative abundance of sympatric dasyurid carnivores: A hypothesis of competition. J. Anim. Ecol. 67, 410–421 (1998).Article 

    Google Scholar 
    Pemberton, D. et al. The diet of the Tasmanian Devil, Sarcophilus harrisii, as determined from analysis of scat and stomach contents. Pap. Proc. R. Soc. Tasman. 142, 13–22 (2008).
    Google Scholar 
    Rogers, T. L., Fox, S., Pemberton, D. & Wise, P. Sympathy for the devil: Captive-management style did not influence survival, body-mass change or diet of Tasmanian devils 1 year after wild release. Wildl. Res. 43, 544–552 (2016).Article 

    Google Scholar 
    Andersen, G. E., Johnson, C. N., Barmuta, L. A. & Jones, M. E. Dietary partitioning of Australia’s two marsupial hypercarnivores, the Tasmanian devil and the spotted-tailed quoll, across their shared distributional range. PLoS ONE 12, e0188529 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Department of Primary Industries Parks Water and Environment. Recovery Plan for the Tasmanian devil (Sarcophilus harrisii) (2010).Brown, O. J. F. Tasmanian devil (Sarcophilus harrisii) extinction on the Australian mainland in the mid-Holocene: multicausality and ENSO intensification. Alcheringa Aust. J. Palaeontol. 30, 49–57 (2006).Article 

    Google Scholar 
    Lewis, A. C., Hughes, C. & Rogers, T. L. Effects of intraspecific competition and body mass on diet specialization in a mammalian scavenger. Ecol. Evol. 12, e8338 (2022).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Andersen, G. E., McGregor, H. W., Johnson, C. N. & Jones, M. E. Activity and social interactions in a wide-ranging specialist scavenger, the Tasmanian devil (Sarcophilus harrisii), revealed by animal-borne video collars. PLoS ONE 15, e0230216 (2020).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Jones, M. E. Road upgrade, road mortality and remedial measures: Impacts on a population of eastern quolls and Tasmanian devils. Wildl. Res. 27, 289–296 (2000).Article 

    Google Scholar 
    Jones, M. E. & Barmuta, L. A. Niche differentiation among sympatric australian dasyurid carnivores. J. Mammal. 81, 434–447 (2000).Article 

    Google Scholar 
    Andersen, G. E., Johnson, C. N., Barmuta, L. A. & Jones, M. E. Use of anthropogenic linear features by two medium-sized carnivores in reserved and agricultural landscapes. Sci. Rep. 7, 11624 (2017).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hamede, R. K., McCallum, H. & Jones, M. Seasonal, demographic and density-related patterns of contact between Tasmanian devils (Sarcophilus harrisii): Implications for transmission of devil facial tumour disease. Austral. Ecol. 33, 614–622 (2008).Article 

    Google Scholar 
    Kitchener, A. & Harris, S. From Forest to Fjaeldmark: Descriptions of Tasmania’s Vegetation (Department of Primary Industries, Parks, Water and Environment, Tasmania, 2013).
    Google Scholar 
    Wiggins, N. L. & Bowman, D. M. J. S. Macropod habitat use and response to management interventions in an agricultural—Forest mosaic in north-eastern Tasmania as inferred by scat surveys. Wildl. Res. 38, 103–113 (2011).Article 

    Google Scholar 
    Hobday, A. J. & Minstrell, M. L. Distribution and abundance of roadkill on Tasmanian highways: Human management options. Wildl. Res. 35, 712–726 (2008).Article 

    Google Scholar 
    Hingston, A. B. Impacts of logging on autumn bird populations in the southern forests of Tasmania. Pap. Proc. R. Soc. Tasman. 134, 19–28 (2000).
    Google Scholar 
    Taylor, R. J. Notes on the diet of the carnivorous mammals of the Upper Henty River Region, western Tasmania. Pap. Proc. R. Soc. Tasman. 120, 7–10 (1986).
    Google Scholar 
    Hall-Aspland, S., Rogers, T., Canfield, R. & Tripovich, J. Food transit times in captive leopard seals (Hydrurga leptonyx). Polar Biol. 34, 95–99 (2011).Article 

    Google Scholar 
    Bell, O. et al. Age-related variation in the trophic characteristics of a marsupial carnivore, the Tasmanian devil Sarcophilus harrisii. Ecol. Evol. 10, 7861–7871 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bell, O. et al. Isotopic niche variation in Tasmanian devils Sarcophilus harrisii with progression of devil facial tumor disease. Ecol. Evol. 11, 8038–8053 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bearhop, S., Adams, C. E., Waldron, S., Fuller, R. A. & MacLeod, H. Determining trophic niche width: A novel approach using stable isotope analysis. J. Anim. Ecol. 73, 1007–1012 (2004).Article 

    Google Scholar 
    Layman, C. A. et al. Applying stable isotopes to examine food-web structure: An overview of analytical tools. Biol. Rev. 87, 545–562 (2012).Article 
    PubMed 

    Google Scholar 
    Crawford, K., McDonald, R. A. & Bearhop, S. Applications of stable isotope techniques to the ecology of mammals. Mamm. Rev. 38, 87–107 (2008).Article 

    Google Scholar 
    Bender, M. M., Rouhani, I., Vines, H. M. & Black, C. C. Jr. 13C/12C ratio changes in crassulacean acid metabolism plants. Plant Physiol. 52, 427–430 (1973).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    O’Leary, M. H. Carbon isotope fractionation in plants. Phytochemistry 20, 553–567 (1981).Article 

    Google Scholar 
    Farquhar, G. D., O’Leary, M. H. & Berry, J. A. On the relationship between carbon isotope discrimination and the intercellular carbon dioxide concentration in leaves. Aust. J. Plant Physiol. 9, 121–137 (1982).CAS 

    Google Scholar 
    Cernusak, L. A. et al. Environmental and physiological determinants of carbon isotope discrimination in terrestrial plants. New Phytol. 200, 950–965 (2013).Article 
    CAS 
    PubMed 

    Google Scholar 
    NSW Parliamentary Counsel. Animal Research Act 1985 (NSW Parliamentary Counsel, 1985).
    Google Scholar 
    National Health and Medical Research Council (Australia). Australian Code for the Care and Use of Animals for Scientific Purposes (National Health and Medical Research Council, 2013).
    Google Scholar 
    du Sert, N. P. et al. Reporting animal research: Explanation and elaboration for the ARRIVE guidelines 2.0. PLoS Biol. 18, e3000411 (2020).Article 

    Google Scholar 
    Environmental Systems Research Institute. ArcGIS Desktop Version 10.8.1. https://www.esri.com/en-us/arcgis/products/arcgis-desktop/overview (2020).Tasmanian Vegetation Monitoring and Mapping Program. TASVEG 4.0. Natural Values Conservation Branch, Department of Primary Industries, Parks, Water and Environment thelist.tas.gov.au/app/content/data/geo-meta-data-record?detailRecordUID=b5c7a079-14bc-4b3c-af73-db7585d34cdd (2020).Land Tasmania. LIST Land Tenure. Land Tasmania thelist.tas.gov.au/app/content/data/geo-meta-data-record?detailRecordUID=9b8bf099-d668–433d-981b-a0f8f964f827 (2015).Hickey, J. E. & Wilkinson, G. R. The development and current implementation of silvicultural pratices in native forests in Tasmania. Aust. For. 62, 245–254 (1999).Article 

    Google Scholar 
    Whiteley, S. B. Calculating the sustainable yield of Tasmania’s State forests. Tasforests 11, 23–34 (1999).
    Google Scholar 
    Pemberton, D. Social Organisation and Behaviour of the Tasmanian devil, Sarcophilus harrisii (University of Tasmania, 1990).
    Google Scholar 
    Attard, M. R. G., Lewis, A. C., Wroe, S., Hughes, C. & Rogers, T. L. Whisker growth in Tasmanian devils (Sarcophilus harrisii) and applications for stable isotope studies. Ecosphere 12, e03846 (2021).Article 

    Google Scholar 
    von Bertalanffy, L. Quantitative laws in metabolism and growth. Q. Rev. Biol. 32, 217–231 (1957).Article 

    Google Scholar 
    Rogers, T. L., Fung, J., Slip, D., Steindler, L. & O’Connell, T. C. Calibrating the time span of longitudinal biomarkers in vertebrate tissues when fine-scale growth records are unavailable. Ecosphere 7, e01449 (2016).Article 

    Google Scholar 
    Qi, H., Coplen, T. B., Geilmann, H., Brand, W. A. & Böhlke, J. K. Two new organic reference materials for δ13C and δ15N measurements and a new value for the δ13C of NBS 22 oil. Rapid Commun. Mass Spectrom. 17, 2483–2487 (2003).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Qi, H. et al. A new organic reference material, l-glutamic acid, USGS41a, for δ13C and δ15N measurements—A replacement for USGS41. Rapid Commun. Mass Spectrom. 30, 859–866 (2016).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Bond, A. L. & Hobson, K. A. Reporting stable-isotope ratios in ecology: Recommended terminology. Guidel. Best Pract. Waterbirds 35, 324–331 (2012).
    Google Scholar 
    O’Connell, T. C. & Hedges, R. E. M. Investigations into the effect of diet on modern human hair isotopic values. Am. J. Phys. Anthropol. 108, 409–425 (1999).Article 
    PubMed 

    Google Scholar 
    Jackson, A. L., Inger, R., Parnell, A. C. & Bearhop, S. Comparing isotopic niche widths among and within communities: SIBER—Stable Isotope Bayesian Ellipses in R. J. Anim. Ecol. 80, 595–602 (2011).Article 
    PubMed 

    Google Scholar 
    R Core Team. R: A Language and Environment for Statistical Computing Version 4.2.0. https://www.r-project.org/ (2022).Bartoń, K. MuMIn: Multi-model inference. R Package Version 1.47.1. https://cran.r-project.org/package=MuMIn (2022).Burnham, K. P. & Anderson, D. R. Model Selection and Multimodel Inference: A Practical Information-Theoretic Approach (Colorado Cooperative Fish and Wildlife Research Unit, 2002).MATH 

    Google Scholar 
    Stock, B. C. et al. Analyzing mixing systems using a new generation of Bayesian tracer mixing models. PeerJ 6, e5096 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Stock, B. C. & Semmens, B. X. MixSIAR: Bayesian Mixing Models in R. R Package Version 3.1.12. https://doi.org/10.5281/zenodo.1209993 (2022).Plummer, M., Stukalov, A. & Denwood, M. rjags: Bayesian graphical models using MCMC. R Package Version 4-13. https://cran.r-project.org/web/packages/rjags/rjags.pdf (2022).Newsome, S. D. et al. Variation in δ13C and δ15N diet–vibrissae trophic discrimination factors in a wild population of California sea otters. Ecol. Appl. 20, 1744–1752 (2010).Article 
    PubMed 

    Google Scholar 
    Brooks, T. M. et al. Habitat loss and extinction in the hotspots of biodiversity. Conserv. Biol. 16, 909–923 (2002).Article 

    Google Scholar 
    Fahrig, L. Effects of habitat fragmentation on biodiversity. Annu. Rev. Ecol. Evol. Syst. 34, 487–515 (2003).Article 

    Google Scholar 
    Pardini, R., Nichols, E. & Püttker, T. Biodiversity response to habitat loss and fragmentation. Encycl. Anthr. 3, 229–239 (2018).Article 

    Google Scholar 
    Koch, A., Munks, S. & Driscoll, D. The use of hollow-bearing trees by vertebrate fauna in wet and dry Eucalyptus obliqua forest, Tasmania. Wildl. Res. 35, 727–746 (2008).Article 

    Google Scholar 
    Donázar, J. A., Cortés-Avizanda, A. & Carrete, M. Dietary shifts in two vultures after the demise of supplementary feeding stations: consequences of the EU sanitary legislation. Eur. J. Wildl. Res. 56, 613–621 (2010).Article 

    Google Scholar 
    Carbone, C., Teacher, A. & Rowcliffe, J. M. The costs of carnivory. PLoS Biol. 5, e22 (2007).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Tucker, M. A., Ord, T. J. & Rogers, T. L. Revisiting the cost of carnivory in mammals. J. Evol. Biol. 29, 2181–2190 (2016).Article 
    CAS 
    PubMed 

    Google Scholar 
    Carbone, C., Mace, G. M., Roberts, S. C. & Macdonald, D. W. Energetic constraints on the diet of terrestrial carnivores. Nature 402, 286–288 (1999).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Fisher, D. O. & Dickman, C. R. Body size-prey relationships in insectivorous marsupials: Tests of three hypotheses. Ecology 74, 1871–1883 (1993).Article 

    Google Scholar 
    Ruxton, G. D. & Houston, D. C. Obligate vertebrate scavengers must be large soaring fliers. J. Theor. Biol. 228, 431–436 (2004).Article 
    ADS 
    MathSciNet 
    PubMed 
    MATH 

    Google Scholar 
    Pemberton, D. & Renouf, D. A field-study of communication and social-behavior of the Tasmanian devil at feeding sites. Aust. J. Zool. 41, 507–526 (1993).Article 

    Google Scholar 
    Pye, R. J. et al. A second transmissible cancer in Tasmanian devils. Proc. Natl. Acad. Sci. USA 113, 374–379 (2016).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    James, S. et al. Tracing the rise of malignant cell lines: Distribution, epidemiology and evolutionary interactions of two transmissible cancers in Tasmanian devils. Evol. Appl. 12, 1772–1780 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hawkins, C. E. et al. Emerging disease and population decline of an island endemic, the Tasmanian devil Sarcophilus harrisii. Biol. Conserv. 131, 307–324 (2006).Article 

    Google Scholar 
    Pearse, A.-M. & Swift, K. Transmission of devil facial-tumour disease. Nature 439, 549 (2006).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Wood, S. W., Hua, Q. & Bowman, D. M. J. S. Fire-patterned vegetation and the development of organic soils in the lowland vegetation mosaics of south-west Tasmania. Aust. J. Bot. 59, 126–136 (2011).Article 

    Google Scholar 
    Kohn, M. J. Carbon isotope compositions of terrestrial C3 plants as indicators of (paleo)ecology and (paleo)climate. PNAS 107, 19691–19695 (2010).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Mayer, M., Ullmann, W., Sunde, P., Fischer, C. & Blaum, N. Habitat selection by the European hare in arable landscapes: The importance of small-scale habitat structure for conservation. Ecol. Evol. 8, 11619–11633 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Barker, R. & Vestjens, W. Food of Australian Birds 1. Non-Passerines (CSIRO Publishing, 1989).Book 

    Google Scholar 
    Thomas, D. G. The bird community of Tasmanian temperate rainforest. Ibis 122, 298–306 (1980).Article 

    Google Scholar 
    DeVault, T. L., Rhodes, O. E. Jr. & Shivik, J. A. Scavenging by vertebrates: Behavioral, ecological, and evolutionary perspectives on an important energy transfer pathway in terrestrial ecosystems. Oikos 102, 225–234 (2003).Article 

    Google Scholar 
    DPIPWE. Annual Statewide Spotlight Surveys, Tasmania 2020/2021. Nature Conservation Report 21/2. (2021).Nguyen, H. K. D., Fielding, M. W., Buettel, J. C. & Brook, B. W. Habitat suitability, live abundance and their link to road mortality of Tasmanian wildlife. Wildl. Res. 46, 236–246 (2019).Article 

    Google Scholar  More

  • in

    Machine learning identifies straightforward early warning rules for human Puumala hantavirus outbreaks

    We performed data acquisition, processing, analysis and visualization using Python23 version 3.8 with the packages Numpy24, Pandas25, Geopandas26, Matplotlib27, Selenium, Beautiful Soup28, SciPy14 and scikit-learn29. The functions used for specific tasks are explicitly mentioned to allow validation and replication studies.Data acquisition and processingHuman PUUV-incidenceHantavirus disease has been notifiable in Germany since 2001. The Robert Koch Institute collects anonymized data from the local and state public health departments and offers via the SurvStat application2 a freely available, limited version of its database for research and informative purposes. We retrieved the reported laboratory-confirmed human PUUV-infections (({text{n}}=text{11,228}) from 2006 to 2021, status: 2022-02-07). From the attributes available for each case, we retrieved the finest temporal and spatial resolution, i.e., the week and the year of notification, together with the district (named “County” in the English version of the SurvStat interface).To avoid bias through underreporting, our dataset was limited to PUUV-infections since 2006. The years 2006–2021 contain 91.9% of the total cases from 2001 to 2021. Human PUUV-incidence was calculated as the number of infections per 100,000 people, by using population data from Eurostat30. For each year, we used the population reported for the January 1 of that year. The population for 2020 was also used for 2021.In the analysis, we only included districts where the total infections were (ge {20}) and the maximum annual incidence was (ge {2}) in the period 2006–2021. The spatial information about the infections provided by the SurvStat application refers to the district where the infection was reported. Therefore, in most of the cases, the reported district corresponds to the residence of the infected person, which may differ from the district of infection. To compensate partially for differences between the reported place of residence and the place of infection, we combined most of the urban districts with their surrounding rural district. The underlying assumption was that most infections reported in urban districts occurred in the neighboring or surrounding rural district. In addition, some urban and rural districts have the same health department. Supplementary Table 1 lists the combined districts.Weather dataFrom the German Meteorological Service31 we retrieved grids of the following monthly weather parameters over Germany from 2004 to 2021: mean daily air temperature—Tmean, minimum daily air temperature—Tmin, and maximum daily air temperature—Tmax (all temperatures are the monthly averages of the corresponding daily values, in 2 m height above ground, in °C); total precipitation in mm—Pr, total sunshine duration in hours—SD, mean monthly soil temperature in 5 cm depth under uncovered typical soil of location in °C—ST, and soil moisture under grass and sandy loam in percent plant useable water—SM. The dataset version for Tmean, Tmin, Tmax, Pr, and SD was v1.0; for ST and SM the dataset version was 0. × . The spatial resolution was 1 × 1 km2.The data acquisition was performed with the Selenium package. The processing was based on the geopandas package26 using a geospatial vector layer for the district boundaries of Germany32. Each grid was processed to obtain the average value of the parameter over each district. We first used the function within to define a mask based on the grid centers contained in the district; we then applied this mask to the grid. In this method, called “central point rasterizing”33, each rectangle of the grid was assigned to a single district, the one that contained its center. The typical processing error was estimated to be about 1%, which agrees with the rasterizing error reported by Bregt et al.33; we consider that most likely this error is significantly less than the uncertainties of the grids themselves, caused by calculation, interpolation, and erroneous or missing observations.Data structureOur analysis was performed at the district level based on the annual infections, acquired by aggregating the weekly cases. From each monthly weather parameter, we created 24 records, for all months of the two previous years. Each observation in our dataset characterized one district in one year. Its target was acquired by transforming the annual incidence, as described in the following section. Each observation comprised all 168 available predictors from the weather parameters (7 parameters × 24 months), thereafter called “variables”. The notation for the naming of the variables follows the format Vx__, where “Vx” can be V1 or V2 that corresponds to one or two years before, respectively;  is the abbreviation of the weather parameter (see previous subsection: “Weather data”); and  is the numerical value of the month, i.e., from 1 to 12.The observations for combined districts retained the label of the rural district. For their infections and populations, we aggregated the individual values, and recalculated the incidence. For their weather variables, we assigned the mean values weighted by the area of each district.Target transformationTo consider the effects that drive the occurrence of high district-relative incidence, we discretized the incidence at the district level. The incidence scaled at its maximum value for each district showed extreme values for minima and maxima. About 49% of all observations were in the range [0, 0.1) and 8% in the range [0.9, 1] (Fig. 5). Therefore, we specifically selected to discretize the scaled incidence with two bins, i.e., to binarize it.Figure 5Histograms of the annual PUUV incidence from 2006 to 2021, scaled to its maximum value for each of the selected districts. Left: Raw incidence. Right: Log-transformed incidence, according to Eq. (6).Full size imageWe first applied a log-transformation to the incidence values34, described in Eq. (6).$${text{Log – incidence}} = log_{10} left( {{text{incidence}} + 1} right)$$
    (6)
    The addition of a positive constant ensured a noninfinite value for zero incidence, with 1 selected so that the log-incidence is nonnegative, and a zero incidence was transformed into a zero log-incidence. This transformation aimed to increase the influence of nonzero incidence values; values that are not pronounced, but still hint at a nonzero infection risk. Its effect is demonstrated in the right plot of Fig. 5, where the positive skewness of the original data is reduced, i.e., low incidence values are spread to higher values, resulting to more uniform bin heights in the range [0.05, 0.95] after the transformation. Formally, in this case the log-transformation achieves a more uniform distribution for the non-extreme incidence values.For the binarization, we performed unsupervised clustering of the log-transformed incidence, separately for each district, applying the function KBinsDiscretizer of the scikit-learn package29. Our selected strategy was the k-means clustering with two bins, because it does not require a pre-defined threshold, and it can operate with the same fixed number of bins for every district, by automatically adjusting the cluster centroids accordingly.Classification methodWe concentrated only on those variable combinations that led to a linear decision boundary for the classification of our selected target. We selected support vector machines (SVM)35 with a linear kernel, because they combine high performance with low model complexity, in that they return the decision boundary as a linear equation of the variables. In addition, SVM is geometrically motivated36 and expected to be less prone to outliers and overfitting than other machine-learning classification algorithms, such as the logistic regression. For the complete modelling process, the regularization parameter C was set to 1, that is the default value in the applied SVC method of the scikit-learn package29, and the weights for both risk classes were also set to 1.Feature selectionOur aim was to use the smallest possible number of weather parameters as variables for a classification model with sufficient performance. To identify the optimal variable combination, we first applied an SVM with a linear kernel for all 2-variable combinations of the monthly weather variables from V2 and V1, i.e., 168 variables (7 weather parameters × 2 years × 12 months). Only for this step, the variables were scaled to their minimum and maximum values, which significantly reduced the processing time. For all the following steps, the scaler was omitted, because the unscaled support vectors were required for the final model. From the total 14,028 models for each unique pair ((frac{168!}{2!cdot left(168-2right)!})), we kept the 100 models with the best F1-score, i.e., of the harmonic mean of sensitivity and precision, and counted the occurrences of each year-month combination in the variables. The best F1-score was 0.752 for the pair (V1_Tmean_9 and V2_Tmax_4); and the best sensitivity was 83% for the pair (V2_Tmax_9 and V1_ST_9).The year-month combinations with more than 10% occurrences were: V1_9 (September of the previous year, with 49% occurrences), V2_9 (September of two years before, with 12%) and V2_4 (April of two years before, with 10%). To avoid sets with highly correlated variables, we formed 3-variable combinations, with exactly one variable from each year-month combination (threefold Cartesian product). From the total 343 models (73 combinations, i.e., 7 weather parameters for 3 year-month combinations), we selected the model with the best sensitivity and at least 70% precision, i.e., the variable set (V2_ST_4, V2_SD_9, and V1_ST_9). We consider that the criteria for this selection are not particularly crucial; and we expect comparable performance for most variable sets with a high F1-score, because the variables for each dimension of the Cartesian product were highly correlated. The eight variable sets with at least 70% precision and at least 80% sensitivity are shown in Supplementary Table 2.The SVM classifier has two hyperparameters: the regularization parameter C and the class weights. By decreasing C, the decision boundary becomes softer and more misclassifications are allowed. On the other hand, increasing the high-risk class weight, the misclassifications of high-risk observations are penalized higher, which is expected to increase the sensitivity and decrease the precision. The simultaneous adjustment of both hyperparameters ensures that the resulting model has the optimal performance with respect to the preferred metric. However, in order to avoid overfitting, we considered redundant a further model optimization with these two hyperparameters. For completeness, we examined SVM models for different values of the hyperparameters and found that the global maximum for the F1-score is in the region of 0.001 for C and 1.5 for the high-risk class weight. Our selected values C = 1 and high-risk class weight equal to 1 give the second best F1-score, which is a local maximum with comparable performance, mostly insensitive to the selection of C from the range [0.2, 5.5].The addition of a fourth variable from V1_6 (June of the previous year) resulted in a model with higher sensitivity but lower precision and specificity (for V1_Pr_6). The highest F1-score was achieved for the quadruple (V2_ST_4, V2_SD_9, V1_ST_9, V1_Pr_6). Because of the increased complexity without significant improvement in the performance, we considered unnecessary a further expansion of our variable triplet. More