More stories

  • in

    Behaviour dominates impacts

    The impacts of climate change on host–parasite dynamics are particularly complex to predict, as they involve an interplay of both physiological and behavioural factors, from both host and parasite. For example, while warming may increase parasite developmental rates and thus increase transmission, excessive heat may instead exceed thermal limits, leading to higher parasite mortality. Transmission also relates to both the distribution and abundance of host species, which may also shift under changing climates. More

  • in

    Climate change impacts the vertical structure of marine ecosystem thermal ranges

    Barnett, T. P. et al. Penetration of human-induced warming into the world’s oceans. Science 309, 284–287 (2005).CAS 
    Article 

    Google Scholar 
    Levitus, S. et al. Global ocean heat content 1955–2008 in light of recently revealed instrumentation problems. Geophys. Res. Lett. 36, L07608 (2009).
    Google Scholar 
    Poloczanska, E. S. et al. Global imprint of climate change on marine life. Nat. Clim. Change 3, 919–925 (2013).Article 

    Google Scholar 
    García Molinos, J. et al. Climate velocity and the future global redistribution of marine biodiversity. Nat. Clim. Change 6, 83–88 (2016).Article 

    Google Scholar 
    Free, C. M. et al. Impacts of historical warming on marine fisheries production. Science 363, 979–983 (2019).CAS 
    Article 

    Google Scholar 
    Hughes, N. F. & Grand, T. C. Physiological ecology meets the ideal-free distribution: predicting the distribution of size-structured fish populations across temperature gradients. Environ. Biol. Fishes 59, 285–298 (2000).Article 

    Google Scholar 
    Tittensor, D. P. et al. Global patterns and predictors of marine biodiversity across taxa. Nature 466, 1098–1101 (2010).CAS 
    Article 

    Google Scholar 
    Sunday, J. M., Bates, A. E. & Dulvy, N. K. Global analysis of thermal tolerance and latitude in ectotherms. Proc. R. Soc. B 278, 1823–1830 (2011).Article 

    Google Scholar 
    Waldock, C., Stuart‐Smith, R. D., Edgar, G. J., Bird, T. J. & Bates, A. E. The shape of abundance distributions across temperature gradients in reef fishes. Ecol. Lett. 22, 685–696 (2019).Article 

    Google Scholar 
    Stuart-Smith, R. D., Edgar, G. J. & Bates, A. E. Thermal limits to the geographic distributions of shallow-water marine species. Nat. Ecol. Evol. 1, 1846–1852 (2017).Article 

    Google Scholar 
    Pinsky, M. L., Worm, B., Fogarty, M. J., Sarmiento, J. L. & Levin, S. A. Marine taxa track local climate velocities. Science 341, 1239–1242 (2013).CAS 
    Article 

    Google Scholar 
    Beaugrand, G., Edwards, M., Raybaud, V., Goberville, E. & Kirby, R. R. Future vulnerability of marine biodiversity compared with contemporary and past changes. Nat. Clim. Change 5, 695–701 (2015).Article 

    Google Scholar 
    Trisos, C. H., Merow, C. & Pigot, A. L. The projected timing of abrupt ecological disruption from climate change. Nature 580, 496–501 (2020).CAS 
    Article 

    Google Scholar 
    Levin, L. A. & Le Bris, N. The deep ocean under climate change. Science 350, 766–768 (2015).CAS 
    Article 

    Google Scholar 
    Deutsch, C. A. et al. Impacts of climate warming on terrestrial ectotherms across latitude. Proc. Natl Acad. Sci. USA 105, 6668–6672 (2008).CAS 
    Article 

    Google Scholar 
    Sunday, J. M., Bates, A. E. & Dulvy, N. K. Thermal tolerance and the global redistribution of animals. Nat. Clim. Change 2, 686–690 (2012).Article 

    Google Scholar 
    Radeloff, V. C. et al. The rise of novelty in ecosystems. Ecol. Appl. 25, 2051–2068 (2015).Article 

    Google Scholar 
    Lotterhos, K. E., Láruson, Á. J. & Jiang, L.-Q. Novel and disappearing climates in the global surface ocean from 1800 to 2100. Sci. Rep. 11, 15535 (2021).CAS 
    Article 

    Google Scholar 
    Mora, C. et al. The projected timing of climate departure from recent variability. Nature 502, 183–187 (2013).CAS 
    Article 

    Google Scholar 
    Henson, S. A. et al. Rapid emergence of climate change in environmental drivers of marine ecosystems. Nat. Commun. 8, 14682 (2017).Article 

    Google Scholar 
    Séférian, R. et al. Evaluation of CNRM Earth System Model, CNRM‐ESM2‐1: role of Earth system processes in present‐day and future climate. J. Adv. Model. Earth Syst. 11, 4182–4227 (2019).Article 

    Google Scholar 
    Gidden, M. J. et al. Global emissions pathways under different socioeconomic scenarios for use in CMIP6: a dataset of harmonized emissions trajectories through the end of the century. Geosci. Model Dev. 12, 1443–1475 (2019).CAS 
    Article 

    Google Scholar 
    Eyring, V. et al. Overview of the Coupled Model Intercomparison Project Phase 6 (CMIP6) experimental design and organization. Geosci. Model Dev. 9, 1937–1958 (2016).Article 

    Google Scholar 
    Beszczynska-Möller, A., Fahrbach, E., Schauer, U. & Hansen, E. Variability in Atlantic water temperature and transport at the entrance to the Arctic Ocean, 1997–2010. ICES J. Mar. Sci. 69, 852–863 (2012).Article 

    Google Scholar 
    Sutton, T. T. Vertical ecology of the pelagic ocean: classical patterns and new perspectives. J. Fish. Biol. 83, 1508–1527 (2013).CAS 
    Article 

    Google Scholar 
    Richter, I. Climate model biases in the eastern tropical oceans: causes, impacts and ways forward. WIREs Clim. Change 6, 345–358 (2015).Article 

    Google Scholar 
    Pozo Buil, M. et al. A dynamically downscaled ensemble of future projections for the California Current System. Front. Mar. Sci. 8, 612874 (2021).Article 

    Google Scholar 
    Leonard, M. et al. A compound event framework for understanding extreme impacts. WIREs Clim. Change 5, 113–128 (2014).Article 

    Google Scholar 
    Kwiatkowski, L. et al. Twenty-first century ocean warming, acidification, deoxygenation, and upper-ocean nutrient and primary production decline from CMIP6 model projections. Biogeosciences 17, 3439–3470 (2020).CAS 
    Article 

    Google Scholar 
    Bopp, L. et al. Multiple stressors of ocean ecosystems in the 21st century: projections with CMIP5 models. Biogeosciences 10, 6225–6245 (2013).Article 

    Google Scholar 
    Cheng, L., Abraham, J., Hausfather, Z. & Trenberth, K. E. How fast are the oceans warming? Science 363, 128–129 (2019).CAS 
    Article 

    Google Scholar 
    Hawkins, E. & Sutton, R. Time of emergence of climate signals. Geophys. Res. Lett. 39, L01702 (2012).Article 

    Google Scholar 
    Stuart-Smith, R. D., Edgar, G. J., Barrett, N. S., Kininmonth, S. J. & Bates, A. E. Thermal biases and vulnerability to warming in the world’s marine fauna. Nature 528, 88–92 (2015).CAS 
    Article 

    Google Scholar 
    Filbee-Dexter, K. et al. Marine heatwaves and the collapse of marginal North Atlantic kelp forests. Sci. Rep. 10, 13388 (2020).CAS 
    Article 

    Google Scholar 
    Román-Palacios, C. & Wiens, J. J. Recent responses to climate change reveal the drivers of species extinction and survival. Proc. Natl Acad. Sci. USA 117, 4211–4217 (2020).Article 
    CAS 

    Google Scholar 
    Silvy, Y., Guilyardi, E., Sallée, J.-B. & Durack, P. J. Human-induced changes to the global ocean water masses and their time of emergence. Nat. Clim. Change 10, 1030–1036 (2020).CAS 
    Article 

    Google Scholar 
    Cheng, L., Zheng, F. & Zhu, J. Distinctive ocean interior changes during the recent warming slowdown. Sci. Rep. 5, 14346 (2015).CAS 
    Article 

    Google Scholar 
    Brito-Morales, I. et al. Climate velocity reveals increasing exposure of deep-ocean biodiversity to future warming. Nat. Clim. Change 10, 576–581 (2020).CAS 
    Article 

    Google Scholar 
    Frölicher, T. L. & Laufkötter, C. Emerging risks from marine heat waves. Nat. Commun. 9, 650 (2018).Article 
    CAS 

    Google Scholar 
    Oliver, E. C. J. et al. Marine Heatwaves. Ann. Rev. Mar. Sci. 13, 313–342 (2021).Article 

    Google Scholar 
    Perry, A. L., Low, P. J., Ellis, J. R. & Reynolds, J. D. Climate change and distribution shifts in marine fishes. Science 308, 1912–1915 (2005).CAS 
    Article 

    Google Scholar 
    Chaudhary, C., Richardson, A. J., Schoeman, D. S. & Costello, M. J. Global warming is causing a more pronounced dip in marine species richness around the equator. Proc. Natl Acad. Sci. USA 118, e2015094118 (2021).CAS 
    Article 

    Google Scholar 
    Burrows, M. T. et al. Ocean community warming responses explained by thermal affinities and temperature gradients. Nat. Clim. Change 9, 959–963 (2019).Article 

    Google Scholar 
    IPCC Climate Change 2022: Impacts, Adaptation, and Vulnerability (eds Pörtner, H.-O. et al.) (Cambridge Univ. Press, 2022).Cahill, A. E. et al. How does climate change cause extinction? Proc. R. Soc. B280, 20121890 (2013).Article 

    Google Scholar 
    Hastings, R. A. et al. Climate change drives poleward increases and equatorward declines in marine species. Curr. Biol. 30, 1572–1577.e2 (2020).CAS 
    Article 

    Google Scholar 
    Jorda, G. et al. Ocean warming compresses the three-dimensional habitat of marine life. Nat. Ecol. Evol. 4, 109–114 (2020).Article 

    Google Scholar 
    Dulvy, N. K. et al. Climate change and deepening of the North Sea fish assemblage: a biotic indicator of warming seas. J. Appl. Ecol. 45, 1029–1039 (2008).Article 

    Google Scholar 
    Thatje, S. Climate warming affects the depth distribution of marine ectotherms. Mar. Ecol. Prog. Ser. 660, 233–240 (2021).Article 

    Google Scholar 
    Manuel, S. A., Coates, K. A., Kenworthy, W. J. & Fourqurean, J. W. Tropical species at the northern limit of their range: composition and distribution in Bermuda’s benthic habitats in relation to depth and light availability. Mar. Environ. Res. 89, 63–75 (2013).CAS 
    Article 

    Google Scholar 
    Peck, L. S., Webb, K. E. & Bailey, D. M. Extreme sensitivity of biological function to temperature in Antarctic marine species. Funct. Ecol. 18, 625–630 (2004).Article 

    Google Scholar 
    Peck, L. S., Morley, S. A., Richard, J. & Clark, M. S. Acclimation and thermal tolerance in Antarctic marine ectotherms. J. Exp. Biol. 217, 16–22 (2014).Article 

    Google Scholar 
    Walsh, J. E. Climate of the Arctic marine environment. Ecol. Appl. 18, S3–S22 (2008).Article 

    Google Scholar 
    Storch, D., Menzel, L., Frickenhaus, S. & Pörtner, H.-O. Climate sensitivity across marine domains of life: limits to evolutionary adaptation shape species interactions. Glob. Change Biol. 20, 3059–3067 (2014).Article 

    Google Scholar 
    Araújo, M. B. et al. Heat freezes niche evolution. Ecol. Lett. 16, 1206–1219 (2013).Article 

    Google Scholar 
    Pörtner, H. O., Peck, L. & Somero, G. Thermal limits and adaptation in marine Antarctic ectotherms: an integrative view. Philos. Trans. R. Soc. B 362, 2233–2258 (2007).Article 
    CAS 

    Google Scholar 
    Qu, Y.-F. & Wiens, J. J. Higher temperatures lower rates of physiological and niche evolution. Proc. R. Soc. B 287, 20200823 (2020).Article 

    Google Scholar 
    Cohen, D.M., Inada, T., Iwamoto, T. and Scialabba, N. FAO Species Catalogue, Vol. 10. Gadiform Fishes of the World (Order Gadiformes) (FAO, 1990).Strand, E. & Huse, G. Vertical migration in adult Atlantic cod (Gadus morhua). Can. J. Fish. Aquat. Sci. 64, 1747–1760 (2007).Article 

    Google Scholar 
    Frölicher, T. L., Fischer, E. M. & Gruber, N. Marine heatwaves under global warming. Nature 560, 360–364 (2018).Article 
    CAS 

    Google Scholar 
    Wernberg, T. et al. Climate-driven regime shift of a temperate marine ecosystem. Science 353, 169–172 (2016).CAS 
    Article 

    Google Scholar 
    Smale, D. A. et al. Marine heatwaves threaten global biodiversity and the provision of ecosystem services. Nat. Clim. Change 9, 306–312 (2019).Article 

    Google Scholar 
    Cheung, W. W. L. & Frölicher, T. L. Marine heatwaves exacerbate climate change impacts for fisheries in the northeast Pacific. Sci. Rep. 10, 6678 (2020).CAS 
    Article 

    Google Scholar 
    Brierley, A. S. & Kingsford, M. J. Impacts of climate change on marine organisms and ecosystems. Curr. Biol. 19, R602–R614 (2009).CAS 
    Article 

    Google Scholar 
    Bijma, J., Pörtner, H.-O., Yesson, C. & Rogers, A. D. Climate change and the oceans—what does the future hold? Mar. Pollut. Bull. 74, 495–505 (2013).CAS 
    Article 

    Google Scholar 
    Jackson, J. B. C. et al. Historical overfishing and the recent collapse of coastal ecosystems. Science 293, 629–637 (2001).CAS 
    Article 

    Google Scholar 
    Duarte, C. M. et al. The soundscape of the Anthropocene ocean. Science 371, eaba4658 (2021).CAS 
    Article 

    Google Scholar 
    Rochman, C. M. & Hoellein, T. The global odyssey of plastic pollution. Science 368, 1184–1185 (2020).CAS 
    Article 

    Google Scholar 
    Gruber, N., Boyd, P. W., Frölicher, T. L. & Vogt, M. Biogeochemical extremes and compound events in the ocean. Nature 600, 395–407 (2021).CAS 
    Article 

    Google Scholar 
    Madec, G. et al. NEMO ocean engine. Zenodo https://www.earth-prints.org/handle/2122/13309 (2017).Mathiot, P., Jenkins, A., Harris, C. & Madec, G. Explicit representation and parametrised impacts of under ice shelf seas in the z∗- coordinate ocean model NEMO 3.6. Geosci. Model Dev. 10, 2849–2874 (2017).Article 

    Google Scholar 
    Dai, A. & Bloecker, C. E. Impacts of internal variability on temperature and precipitation trends in large ensemble simulations by two climate models. Clim. Dyn. 52, 289–306 (2019).Article 

    Google Scholar 
    Deser, C., Phillips, A., Bourdette, V. & Teng, H. Uncertainty in climate change projections: the role of internal variability. Clim. Dyn. 38, 527–546 (2012).Article 

    Google Scholar 
    Middag, R. et al. Intercomparison of dissolved trace elements at the Bermuda Atlantic Time Series station. Mar. Chem. 177, 476–489 (2015).CAS 
    Article 

    Google Scholar 
    Welch, B. L. The generalization of Student’s’ problem when several different population variances are involved. Biometrika 34, 28 (1947).CAS 

    Google Scholar 
    Lenoir, J. et al. Species better track climate warming in the oceans than on land. Nat. Ecol. Evol. 4, 1044–1059 (2020).Article 

    Google Scholar 
    Janzen, D. H. Why mountain passes are higher in the Tropics. Am. Nat. 101, 233–249 (1967).Article 

    Google Scholar 
    Seebacher, F., White, C. R. & Franklin, C. E. Physiological plasticity increases resilience of ectothermic animals to climate change. Nat. Clim. Change 5, 61–66 (2015).Article 

    Google Scholar 
    Hoffmann, A. A. & Sgrò, C. M. Climate change and evolutionary adaptation. Nature 470, 479–485 (2011).CAS 
    Article 

    Google Scholar 
    Sandblom, E. et al. Physiological constraints to climate warming in fish follow principles of plastic floors and concrete ceilings. Nat. Commun. 7, 11447 (2016).CAS 
    Article 

    Google Scholar 
    Tewksbury, J. J., Huey, R. B. & Deutsch, C. A. Putting the heat on tropical animals. Science 320, 1296–1297 (2008).CAS 
    Article 

    Google Scholar 
    Dahlke, F. T., Wohlrab, S., Butzin, M. & Pörtner, H.-O. Thermal bottlenecks in the life cycle define climate vulnerability of fish. Science 369, 65–70 (2020).CAS 
    Article 

    Google Scholar  More

  • in

    Spring thaw nitrous oxide

    Agriculture soils are a source of nitrous oxide and account for 60% of total emissions. It is well established that nitrogen addition via fertilizers drives nitrous oxide emissions during crop growing season. However, little is known about the role of melting snow and thawing surface soil layers during the spring. Limited knowledge of this phenomenon reduces our ability to develop accurate nitrous oxide emissions inventories required under the UN Framework Convention on Climate Change (UNFCCC). More

  • in

    Honey bees save energy in honey processing by dehydrating nectar before returning to the nest

    Berenbaum, M. R. & Calla, B. Honey as a functional food for Apis mellifera. Annu. Rev. Entomol. 66, 185–208. https://doi.org/10.1146/annurev-ento-040320-074933 (2021).CAS 
    Article 
    PubMed 

    Google Scholar 
    Crane, E. Honey: A Comprehensive Survey (Heinemann, 1975).
    Google Scholar 
    Park, O. W. The storing and ripening of honey by honeybees. J. Econ. Entomol. 18, 405–410 (1925).Article 

    Google Scholar 
    Reinhardt, J. F. Ventilating the bee colony to facilitate the honey ripening process. J. Econ. Entomol. 32, 654–660. https://doi.org/10.1093/jee/32.5.654 (1939).Article 

    Google Scholar 
    Eyer, M., Neumann, P. & Dietemann, V. A look into the cell: Honey storage in honey bees, Apis mellifera. PLoS ONE 11(8), e0161059 (2016).Article 

    Google Scholar 
    Oertel, E., Fieger, E. A., Williams, V. R. & Andrews, E. A. Inversion of cane sugar in the honey stomach of the bee. J. Econ. Entomol. 44, 487–492 (1951).CAS 
    Article 

    Google Scholar 
    Park, O. W. Studies on the changes in nectar concentration produced by the honeybee, Apis mellifera. Part I. Changes which occur between the flower and the hive. Res. Bull. Iowa Agric. Exp. Station 151, 211–243 (1932).
    Google Scholar 
    Nicolson, S. W. & Human, H. Bees get a head start on honey production. Biol. Let. 4, 299–301. https://doi.org/10.1098/rsbl.2008.0034 (2008).Article 

    Google Scholar 
    Nicolson, S. W. & Louw, G. N. Simultaneous measurement of evaporative water loss, oxygen consumption, and thoracic temperature during flight in a carpenter bee. J. Exp. Zool. 222, 287–296 (1982).Article 

    Google Scholar 
    Schmid-Hempel, P., Kacelnik, A. & Houston, A. I. Honeybees maximize efficiency by not filling their crop. Behav. Ecol. Sociobiol. 17, 61–66 (1985).Article 

    Google Scholar 
    Kacelnik, A., Houston, A. I. & Schmid-Hempel, P. Central-place foraging in honey bees: The effect of travel time and nectar flow on crop filling. Behav. Ecol. Sociobiol. 19, 19–24. https://doi.org/10.1007/BF00303838 (1986).Article 

    Google Scholar 
    Wolf, T. J., Schmid-Hempel, P., Ellington, C. P. & Stevenson, R. D. Physiological correlates of foraging efforts in honey-bees: Oxygen consumption and nectar load. Funct. Ecol. 3, 417–424 (1989).Article 

    Google Scholar 
    Mitchell, D. Thermal efficiency extends distance and variety for honeybee foragers: Analysis of the energetics of nectar collection and desiccation by Apis mellifera. J. R. Soc. Interface 16, 20180879. https://doi.org/10.1098/rsif.2018.0879 (2019).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Corbet, S. A. et al. Native or exotic? Double or single? Evaluating plants for pollinator-friendly gardens. Ann. Bot. 87, 219–232 (2001).Article 

    Google Scholar 
    Harano, K. & Nakamura, J. Nectar loads as fuel for collecting nectar and pollen in honeybees: Adjustment by sugar concentration. J. Comp. Physiol. A. https://doi.org/10.1007/s00359-016-1088-x (2016).Article 

    Google Scholar 
    Nicolson, S. W. & van Wyk, B.-E. Nectar sugars in Proteaceae: Patterns and processes. Aust. J. Bot. 46, 489–504 (1998).Article 

    Google Scholar 
    Corbet, S. A. Nectar sugar content: Estimating standing crop and secretion rate in the field. Apidologie 34, 1–10. https://doi.org/10.1051/apido:2002049 (2003).CAS 
    Article 

    Google Scholar 
    Southwick, E. E. & Pimentel, D. Energy efficiency of honey production by bees. Bioscience 31, 730–732. https://doi.org/10.2307/1308779 (1981).Article 

    Google Scholar 
    Mitchell, D. Nectar, humidity, honey bees (Apis mellifera) and varroa in summer: A theoretical thermofluid analysis of the fate of water vapour from honey ripening and its implications on the control of Varroa destructor. J. R. Soc. Interface 16, 20190048. https://doi.org/10.1098/rsif.2019.0048 (2019).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Human, H., Nicolson, S. W. & Dietemann, V. Do honeybees, Apis mellifera scutellata, regulate humidity in their nest?. Naturwissenschaften 93, 397–401 (2006).ADS 
    CAS 
    Article 

    Google Scholar 
    Ellis, M. B. Homeostasis: Humidity and water relations in honeybee colonies, MSc thesis, University of Pretoria (2008).Ellis, M., Nicolson, S., Crewe, R. & Dietemann, V. Hygropreference and brood care in the honeybee (Apis mellifera). J. Insect Physiol. 54, 1516–1521. https://doi.org/10.1016/j.jinsphys.2008.08.011 (2008).CAS 
    Article 
    PubMed 

    Google Scholar 
    Portman, Z. M., Ascher, J. S. & Cariveau, D. P. Nectar concentrating behavior by bees (Hymenoptera: Anthophila). Apidologie 52, 1169–1194. https://doi.org/10.1007/s13592-021-00895-1 (2021).Article 

    Google Scholar 
    Nicolson, S. W. Water homeostasis in bees, with the emphasis on sociality. J. Exp. Biol. 212, 429–434. https://doi.org/10.1242/jeb.022343 (2009).Article 
    PubMed 

    Google Scholar 
    Pokorny, T., Lunau, K. & Eltz, T. Raising the sugar content – orchid bees overcome the constraints of suction feeding through manipulation of nectar and pollen provisions. PLoS ONE 9(11), e113823. https://doi.org/10.1371/journal.pone.0113823 (2014).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Lindauer, M. The water economy and temperature regulation of the honeybee colony. Bee World 36, 81–92 (1955).Article 

    Google Scholar 
    Heinrich, B. Mechanisms of body-temperature regulation in honeybees, Apis mellifera. I. Regulation of head temperature. J. Exp. Biol. 85, 61–72 (1980).Article 

    Google Scholar 
    Cooper, P. D., Schaffer, W. M. & Buchmann, S. L. Temperature regulation of honeybees (Apis mellifera) foraging in the Sonoran desert. J. Exp. Biol. 114, 1–15 (1985).Article 

    Google Scholar 
    Louw, G. N. & Hadley, N. F. Water economy of the honeybee: A stoichiometric accounting. J. Exp. Zool. 235, 147–150 (1985).Article 

    Google Scholar 
    Rodney, S. & Purdy, J. Dietary requirements of individual nectar foragers, and colony-level pollen and nectar consumption: A review to support pesticide exposure assessment for honey bees. Apidologie 51, 163–179. https://doi.org/10.1007/s13592-019-00694-9 (2020).Article 

    Google Scholar 
    Drezner-Levy, T., Smith, B. & Shafir, S. The effect of foraging specialization on various learning tasks in the honey bee (Apis mellifera). Behav. Ecol. Sociobiol. 64, 135–148. https://doi.org/10.1007/s00265-009-0829-z (2009).Article 

    Google Scholar 
    Afik, O. & Shafir, S. Effect of ambient temperature on crop loading in the honey bee, Apis mellifera (Hymenoptera: Apidae). Entomologia Generalis 29, 135–148 (2007).Article 

    Google Scholar 
    Seeley, T. D. Honey bee foragers as sensory units of their colonies. Behav. Ecol. Sociobiol. 34, 51–62 (1994).Article 

    Google Scholar 
    Waller, G. D. Evaluating responses of honeybees to sugar solutions using an artificial-flower feeder. Ann. Entomol. Soc. Am. 65, 857–862 (1972).CAS 
    Article 

    Google Scholar 
    Nicolson, S. W., de Veer, L., Köhler, A. & Pirk, C. W. W. Honeybees prefer warmer nectar and less viscous nectar, regardless of sugar concentration. Proc. R. Soc. B: Biol. Sci. 280, 20131597. https://doi.org/10.1098/rspb.2013.1597 (2013).Article 

    Google Scholar 
    Neff, J. L. & Simpson, B. B. The roles of phenology and reward structure in the pollination biology of wild sunflower (Helianthus annuus L., Asteraceae). Israel J. Bot. 39, 197–216 (1990).
    Google Scholar 
    Waller, G. D., Carpenter, E. W. & Ziehl, O. A. Potassium in onion nectar and its probable effect on attractiveness of onion flowers to honey bees. J. Am. Soc. Hortic. Sci. 97, 535–539 (1972).CAS 
    Article 

    Google Scholar 
    Roubik, D. W., Yanega, D., Aluja, M., Buchmann, S. L. & Inouye, D. W. On optimal nectar foraging by some tropical bees (Hymenoptera: Apidae). Apidologie 26, 197–211 (1995).Article 

    Google Scholar 
    Power, E. F., Stabler, D., Borland, A. M., Barnes, J. & Wright, G. A. Analysis of nectar from low-volume flowers: A comparison of collection methods for free amino acids. Methods Ecol. Evol. 9, 734–743. https://doi.org/10.1111/2041-210X.12928 (2018).Article 
    PubMed 

    Google Scholar 
    Pattrick, J. G., Symington, H. A., Federle, W. & Glover, B. J. The mechanics of nectar offloading in the bumblebee Bombus terrestris and implications for optimal concentrations during nectar foraging. J. R. Soc. Interface 17, 20190632. https://doi.org/10.1098/rsif.2019.0632 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Strauss, U., Dietemann, V., Human, H., Crewe, R. M. & Pirk, C. W. W. Resistance rather than tolerance explains survival of savannah honeybees (Apis mellifera scutellata) to infestation by the parasitic mite Varroa destructor. Parasitology 143, 374–387. https://doi.org/10.1017/s0031182015001754 (2016).Article 
    PubMed 

    Google Scholar 
    Dyer, F. C. & Seeley, T. D. Interspecific comparisons of endothermy in honey-bees (Apis): Deviations from the expected size-related patterns. J. Exp. Biol. 127, 1–26. https://doi.org/10.1242/jeb.127.1.1 (1987).Article 

    Google Scholar  More

  • in

    Decomposition stages as a clue for estimating the post-mortem interval in carcasses and providing accurate bird collision rates

    Barrientos, R. et al. A review of searcher efficiency and carcass persistence in infrastructure-driven mortality assessment studies. Biol. Conserv. 222, 146–153 (2018).
    Google Scholar 
    Stevens, B. S., Reese, K. P. & Connelly, J. W. Survival and detectability bias of avian fence collision surveys in sagebrush steppe. J. Wildl. Manag. 75, 437–449 (2011).
    Google Scholar 
    Hunting, K. A Roadmap for PIER Research on Avian Collisions with Power Lines in California. (2002).Barrientos, R. et al. Wire marking results in a small but significant reduction in avian mortality at power lines: A baci designed study. PLoS ONE 7, e32569 (2012).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Costantini, D., Gustin, M., Ferrarini, A. & Dell’Omo, G. Estimates of avian collision with power lines and carcass disappearance across differing environments. Anim. Conserv. 20, 173–181 (2017).
    Google Scholar 
    Jenkins, A. R. et al. Estimating the impacts of power line collisions on Ludwig’s Bustards Neotis ludwigii. Bird Conserv. Int. 21, 303–310 (2011).
    Google Scholar 
    Shaw, J. M., Reid, T. A., Schutgens, M., Jenkins, A. R. & Ryan, P. G. High power line collision mortality of threatened bustards at a regional scale in the Karoo, South Africa. Ibis (Lond. 1859) 1859(160), 431–446 (2018).
    Google Scholar 
    Gómez-Catasús, J. et al. Factors affecting differential underestimates of bird collision fatalities at electric lines: a case study in the Canary Islands. Ardeola 68, 71–94 (2021).
    Google Scholar 
    Ponce, C., Alonso, J. C., Argandoña, G., García Fernández, A. & Carrasco, M. Carcass removal by scavengers and search accuracy affect bird mortality estimates at power lines. Anim. Conserv. 13, 603–612 (2010).
    Google Scholar 
    Bernardino, J. et al. Bird collisions with power lines: State of the art and priority areas for research. Biol. Conserv. 222, 1–13 (2018).
    Google Scholar 
    Brooks, J. W. & Sutton, L. in Veterinary Forensic Pathology (ed. Brooks, J. W.) 43–63 (2018). https://doi.org/10.1007/978-3-319-67172-7_4Brooks, J. W. Postmortem changes in animal carcasses and estimation of the postmortem interval. Vet. Pathol. 53, 929–940 (2016).CAS 
    PubMed 

    Google Scholar 
    Ascensão, F. et al. Beware that the lack of wildlife mortality records can mask a serious impact of linear infrastructures. Glob. Ecol. Conserv. 19, e00661 (2019).
    Google Scholar 
    Hau, T. C., Hamzah, N. H., Lian, H. H. & Amir Hamzah, S. P. A. Decomposition process and post mortem changes: Review. Sains Malaysiana 43, 1873–1882 (2014).
    Google Scholar 
    Cooper, J. E. in Wildlife Forensic Investigation: Principles and Practice (eds. Cooper, J. & Cooper, M.) 237–324 (CRC Press, 2013). https://doi.org/10.1201/b14553Sutherland, A., Myburgh, J., Steyn, M. & Becker, P. J. The effect of body size on the rate of decomposition in a temperate region of South Africa. Forensic Sci. Int. 231, 257–262 (2013).CAS 
    PubMed 

    Google Scholar 
    Valverde, I., Espín, S., María-Mojica, P. & García-Fernández, A. J. Protocol to classify the stages of carcass decomposition and estimate the time of death in small-size raptors. Eur. J. Wildl. Res. 66, 1–13 (2020).
    Google Scholar 
    Goff, M. L. in Current Concepts in Forensic Entomology (eds. Amendt, J., Goff, M., Campobasso, C. & Grassberger, M.) 1–24 (Springer, 2010). https://doi.org/10.1007/978-1-4020-9684-6_1Pittner, S. et al. A field study to evaluate PMI estimation methods for advanced decomposition stages. Int. J. Legal Med. 134, 1361–1373 (2020).PubMed 
    PubMed Central 

    Google Scholar 
    Probst, C. et al. Estimating the postmortem interval of wild boar carcasses. Vet. Sci. 7, 6 (2020).PubMed Central 

    Google Scholar 
    Cambra-Moo, Ó., Delgado-Buscalioni, Á. & Delgado-Buscalioni, R. An approach to the study of variations in early stages of Gallus gallus decomposition. J. Taphon. 6, 21–40 (2008).
    Google Scholar 
    Oates, D., Coggin, J., Hartman, F. & Hoilien, G. Guide to Time of Death in Selected Wildlife Species. (Nebraska Technical Series No. 14. Lincoln, N.E., Nebraska Game and Parks Commission, 1984).Hewadikaram, K. A. & Goff, M. L. Effect of carcass size on rate of decomposition and arthropod succession patterns. Am. J. Forensic Med. Pathol. 12, 240–265 (1991).
    Google Scholar 
    Zhou, C. & Byard, R. W. Factors and processes causing accelerated decomposition in human cadavers—An overview. J. Forensic Leg. Med. 18, 6–9 (2011).PubMed 

    Google Scholar 
    Cockle, D. L. & Bell, L. S. Human decomposition and the reliability of a ‘Universal’ model for post mortem interval estimations. Forensic Sci. Int. 253(136), e1-136.e9 (2015).
    Google Scholar 
    Azevedo, R. R. & Krüger, R. F. The influence of temperature and humidity on abundance and richness of Calliphoridae (Diptera). Iheringia. Série Zool. 103, 145–152 (2013).
    Google Scholar 
    Barnes, K. M. in Wildlife Forensic Investigation: Principles and Practice (eds. Cooper, J. & Cooper, M.) 149–160 (CRC Press, 2013).Mann, R. W., Bass, W. M. & Meadows, L. Time since death and decomposition of the human body: Variables and observations in case and experimental field studies. J. Forensic Sci. 35, 103–111 (1990).CAS 
    PubMed 

    Google Scholar 
    Gliksman, D. et al. Biotic degradation at night, abiotic degradation at day: Positive feedbacks on litter decomposition in drylands. Glob. Change Biol. 23, 1564–1574 (2017).ADS 

    Google Scholar 
    Araujo, P. I., Grasso, A. A., González-Arzac, A., Méndez, M. S. & Austin, A. T. Sunlight and soil biota accelerate decomposition of crop residues in the Argentine Pampas. Agric. Ecosyst. Environ. 330, 107908 (2022).
    Google Scholar 
    Fernández-Palacios, J. M. & Martín-Esquivel, J. L. Naturaleza de las Islas Canarias: Ecología y Conservación. (Turquesa, 2001).Kenward, M. G. & Roger, J. H. An improved approximation to the precision of fixed effects from restricted maximum likelihood. Comput. Stat. Data Anal. 53, 2583–2595 (2009).MathSciNet 
    MATH 

    Google Scholar 
    R Core Team. R: A language and environment for statistical computing. R Foundation for Statistical Computing. http://www.r-project.org (2020).Bates, D., Mächler, M., Bolker, B. M. & Walker, S. C. Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1–48 (2015).
    Google Scholar 
    Kuznetsova, A., Brockhoff, P. B. & Christensen, R. H. B. lmerTest package: Tests in linear mixed effects models. J. Stat. Softw. 82, 1–26 (2017).
    Google Scholar 
    Zeileis, A. & Hothorn, T. Diagnostic checking in regression relationships. R News 2, 7–10 (2002).
    Google Scholar 
    Halekoh, U. & Højsgaard, S. A Kenward–Roger approximation and parametric bootstrap methods for tests in linear mixed models-the R package pbkrtest. J. Stat. Softw. 59, 1–30 (2014).
    Google Scholar 
    Fox, J. & Weisberg, S. An {R} Companion to Applied Regression, Second Edition. (Sage, 2011).Bartoń, K. MuMIn: Multi-Model Inference. (R Package Version 1.43.6, 2019).De Rosario-Martinez, H., Fox, J. & R Core Team. Package ‘phia’ Title Post-Hoc Interaction Analysis. (R Package Version 0.2–1, 2015).Nakagawa, S. & Schielzeth, H. A general and simple method for obtaining R2 from generalized linear mixed-effects models. Methods Ecol. Evol. 4, 133–142 (2013).
    Google Scholar 
    Vass, A. Beyond the grave—Understanding human decomposition. Microbiol. Today 28, 190–192 (2001).
    Google Scholar 
    Gill-King, H. in Forensic Taphonomy: The Postmortem Fate of Human Remains (eds. Haglund, W. D. & Sorg, M. H.) 93–104 (CRC Press, 1996). https://doi.org/10.1201/9781439821923.sec2Campobasso, C. P., Di Vella, G. & Introna, F. Factors affecting decomposition and Diptera colonization. Forensic Sci. Int. 12, 18–27 (2001).
    Google Scholar 
    Austin, A. T., Araujo, P. I. & Leva, P. E. Interaction of position, litter type, and water pulses on decomposition of grasses from the semiarid Patagonian steppe. Ecology 90, 2642–2647 (2009).PubMed 

    Google Scholar 
    Brandt, L. A., Bonnet, C. & King, J. Y. Photochemically induced carbon dioxide production as a mechanism for carbon loss from plant litter in arid ecosystems. J. Geophys. Res. Biogeosci. 114, G02004 (2009).ADS 

    Google Scholar 
    Lee, H., Rahn, T. & Throop, H. An accounting of C-based trace gas release during abiotic plant litter degradation. Glob. Chang. Biol. 18, 1185–1195 (2012).ADS 

    Google Scholar 
    Zepp, R. G., Erickson, D. J., Paul, N. D. & Sulzberger, B. Interactive effects of solar UV radiation and climate change on biogeochemical cycling. Photochem. Photobiol. Sci. 6, 286–300 (2007).CAS 
    PubMed 

    Google Scholar 
    Archer, M. S. Rainfall and temperature effects on the decomposition rate of exposed neonatal remains. Sci. Justice J. Forensic Sci. Soc. 44, 35–41 (2004).Simmons, T., Adlam, R. E. & Moffatt, C. Debugging decomposition data—Comparative taphonomic studies and the influence of insects and carcass size on decomposition rate. J. Forensic Sci. 55, 8–13 (2010).PubMed 

    Google Scholar 
    Spicka, A., Johnson, R., Bushing, J., Higley, L. G. & Carter, D. O. Carcass mass can influence rate of decomposition and release of ninhydrin-reactive nitrogen into gravesoil. Forensic Sci. Int. 209, 80–85 (2011).CAS 
    PubMed 

    Google Scholar 
    Tracqui. in Encyclopaedia of Forensic Sciences (eds. Siegel, J. A., Saukko, P. J. & Max, M. H.) 1357–1363 (Academic Press, 2000).Riding, C. S. & Loss, S. R. Factors influencing experimental estimation of scavenger removal and observer detection in bird–window collision surveys. Ecol. Appl. 28, 2119–2129 (2018).PubMed 

    Google Scholar  More

  • in

    Efficiency of the traditional practice of traps to stimulate black truffle production, and its ecological mechanisms

    Dataset 1: Analysis of truffle growers archivesWe selected eleven T. melanosporum orchards located across the South-West France, from Montpellier (43°44′01.4″N 3°42′13.2″E) to Jonzac (45°27′17.7″N, 0°25′26.9″W; Fig. 2). These sites were selected for (1) the quality of the records of fruitbody production and practices by truffle growers (Table S1), including the detail of inoculations since plantation (amount and frequency of added crushed sporocarps), (2) the use of truffle traps by the owners and the quality of the record from these devices, and (3) the presence of oaks (Quercus ilex, Q. pubescens and Q. suber) as the only hosts tree species. Based on the archives of truffle growers, including a systematic recording of truffle production within and outside traps, we reported at each study site the contribution of truffle traps to the annual fruitbody production of the entire truffle grounds, by using number and/or weight of collected fruitbodies within (Pin) and outside (Pout) truffle traps.Dataset 2: In situ experiment tracing the inoculation effectThree orchards located near Angoulème (45°74′35.5″N, − 0°63′78.4″W), Jonzac (45°44′09.8″N, 0°43′96.7″W), and Arles-sur-Tech (42°45′44.9″N, 2°62′89.4″W), hereafter referred to Site 1 to 3 (Fig. 2) were selected for testing both disturbance effect and inoculum effect on fruitbody production in truffle traps. These sites presented a high fruitbody production and a high Pin/Pout ratio, thus optimum conditions to test mechanisms underlying how truffle traps influence fruitbody production. Host trees were between 5 and 18 years old at the beginning of the experiment (Fig. 2). At each site, we selected three non-adjacent trees (four on Site 3) that displayed a continuous fruitbody production over the three previous years. Under each selected tree, we excavated, at two-thirds of the distance between the tree trunk and the limit of brûlé (a vegetation-poor zone that shows the extension mycelia in the soil40, eight equidistant truffle traps [20 × 20 cm large × 20 cm deep] as shown in Fig. 3a. Under each tree, two traps were filled with only a mixture of peat and vermiculite (hereafter referred as non-inoculated controls) to test for disturbance effect. The used mixture was identical to that which is currently applied in commercial orchards. In three other traps, 5 g of crushed material from a single black truffle fruitbody (including its gleba and spores) were added to the previous mixture (hereafter referred as one mating-type inoculum). In the three last traps, 5 g of crushed material from two ascocarps with gleba of opposite mating types (hereafter referred as two mating-type inoculum) were added to the previous mixture. We added the two mating-type condition to accurately test a potential contribution of the gleba (haploid and thus with a single mating type) on future production. As quoted in Introduction, maternal individuals with opposite mating types tend to exclude each other locally (spatial segregation of clusters of individuals of same mating types26. Thus, the two mating-type inoculum allows us to detect in each trap a maternal contribution by the introduced gleba, despite potential exclusion by pre-installed individuals of the locally dominant mating type in the surrounding. Moreover, it allows us to detect a paternal contribution by the introduced gleba of the mating type opposite to the locally dominant. The eight truffle traps were randomly arranged, so that two repetitions of same modality were always separated by a repetition of another modality (Fig. 3a).In March 2013, six freshly collected truffles (weighting  > 60 g) were molecularly analyzed for the mating type of their gleba as in18. On Site 1 and Site 2, the inoculum was made of fruitbodies collected at Site 1. On Site 3, fruitbodies used as inoculum originated from truffle grounds in Sarrion (Spain). In April 2013, truffles traps were installed as explained above (in all, 8 traps × 3 (or 4) trees × 3 sites) and monitored for two years by truffle growers. Harvesting was performed by trained dogs (one different dog per site) checking truffle traps and the surrounding brûlés at each visit of the orchard by truffle growers. When dogs detected truffles, a small hole was excavated to collect ascocarps without disturbing the trap further. At the end of January, 2015, all truffle traps were completely excavated, remnant truffles overlooked by dogs were systematically collected (Fig. 3b). Three soil aliquots were collected within all traps and pooled. All truffles and soil aliquots were frozen for subsequent DNA analysis.Molecular and genetic analysesDNA extractions, mating typing and genotyping were done as in18. Briefly, DNA was extracted from the gleba and from spores of each fruitbody to get access to the maternal and zygotic DNA, respectively. Simple sequence repeat (SSRs) genotyping was performed using 12 polymorphic markers and the mating-type locus as in18. Gleba extracts displaying apparent heterozygous genotypes, likely due to contamination by spore DNA were systematically discarded from further analyses. For each fruitbody, the haploid paternal genotype was then deduced by subtracting the haploid maternal genotype from the zygotic diploid genotype. This data set was used for relatedness estimations. We discarded from all further analysis the marker me11, which displayed more than 39% missing data, as well as all samples with missing data for at any locus.Multilocus genotypes comparisonsBased on the 11 remaining SSRs and the mating-type (Table S5 and Figure S2), MLGs were identified on all maternal and paternal haploid genomes using GenClone v.2.041, and the probability that MLGs represented more than once resulted from independent events of sexual reproduction was calculated (PSex41,42). On each site, clonal diversity was measured as R = (G − 1)/(N − 1) according to43, where N is the number of fruitbodies and G the number of MLGs. For testing whether the gleba of the inoculated fruitbody contributed, either paternally (H1) or maternally (H2) to the harvested fruitbodies (Fig. 1c), the inoculated maternal MLG was compared to the paternal and maternal MLG of the harvested fruitbodies.Relatedness estimationFor testing whether the spores of the inoculum, which carry many distinct haploid MLGs due to meiosis, had paternal or maternal contribution(s) to the harvested fruitbodies (H3; Fig. 1c), we used relatedness estimation.For testing whether spores of the inoculum had a paternal contribution, an individual relatedness estimate to the spore inoculum was computed for each paternal genome detected in truffle traps. Relatedness r here describes the expected frequency E[p_offpat] of each allele in a given genome, E[p_offpat] = p_pop + r * (p_inoc − p_pop), where p_pop is the allele frequency in the local population (here estimated from the glebas of other truffles collected under the focal tree), and p_inoc is the frequency of the allele in the inoculum. Thus, p_offpat takes values 0 or 1, and p_inoc takes values 0, 0.5 or 1, except when two fruitbodies were used as inoculum (two gleba mating types traps). Thus r = (p_offpat − p_pop)/(p_inoc − p_pop). An individual relatedness estimate for each genome is then obtained by summing over alleles and loci the observed values of the numerator and denominator in this expression. A population-level estimate is further obtained by summing numerators and denominators over the paternity events in each population.To test whether such estimates are compatible with the hypothesis that the paternal individuals are not from the inocula, we obtained the distribution of population-level relatedness estimates by simulating samples under this hypothesis: paternal genotypes were randomly simulated according to alleles frequencies in the local population. For each population, 10,000 samples were simulated, and p-values were estimated as the proportion of simulations with higher population-level relatedness with inocula than the observed one. Confidence intervals for these p-values were computed from the binomial distribution for 10,000 draws, and Bonferroni-corrected over the three populations.For testing whether spores of the inoculum had a maternal contribution (H4, Fig. 1c), we estimated the relatedness of the locally used spore inoculum to each maternal genome detected in truffle traps (deduced from the gleba), and we confronted it to simulated samples as previously but with one modification: if the focal fruitbody was harvested in a trap inoculated with the inoculum A1, all genomes of truffles from traps inoculated with the same inoculum (A1 or A1 + A2 + A3, see Fig. 3c.) were discarded from the estimation of p_pop.Assessment of T. melanosporum mycelium concentration in truffle trapsOn Sites 1, 2 and 3, soil samples were collected in all traps and in the surrounding brûlés at harvesting date (January, 2015). In collected soils, total DNA was extracted and quantified as in19. Briefly, after sieving and homogenizing soil collected in each trap and from out of the brûlés, aliquots (10 g) were analyzed as follows. After extraction with the kit Power Soil (MoBio Laboratories, Carlsbad, CA, USA), the extra-radical mycelium of T. melanosporum was quantified using quantitative Taqman™ PCR (qPCR) with the primers and probe described in44. Triplicate real-time PCR were performed on each sample using the same concentration of primer and the same thermocycling program as in19. Standards were prepared using fresh immature T. melanosporum ascocarp, and a standard curve was generated for each site by plotting serial tenfold dilutions against corresponding initial amount of ascocarp. Absolute quantification of mycelium biomass of T. melanosporum was expressed in mg of mycelium per g of soil.Statistical analysesStatistics were done using R version 4.0.445.Effect of truffle traps on fruitbody production—The contribution of truffle traps to the overall production of orchards was assessed by (1) data mining of truffle growers’ archives (Dataset 1) and (2) comparing the density of truffles harvested in traps (expressed in number of truffles per m2 per orchard; for each sampled tree, traps correspond to an investigated soil surface of s = 8 × 0.2 x 0.2 = 0.32 m2) with the density measured within surrounding brûlés (Dataset 1). On Dataset2, at each site, the area occupied by brûlés was evaluated by measuring in the field the surface of soil devoid of vegetation consecutively to spontaneous T. melanosporum brûlé.Fruitbody production under different conditions (i.e. non-inoculated controls versus one gleba mating type traps versus two gleba mating type traps) were compared using generalized linear mixed models with negative binominal family and log link (R, spam package46). The full model included the logarithm of the sampled area as offset to account for variations in this sampled area, interactions of trap-modality effects with site effect. Formal likelihood ratio tests are based on one-step deletions from this full model, applied to subsets of the data relevant for each hypothesis tested. Additional bootstrap tests (1000 iterations) were run to correct any bias in small sample likelihood ratio tests.Concentrations of T. melanosporum mycelium in soil—Similarly as above, the inoculum effect on mycelium concentrations was compared using generalized linear mixed models with Gamma log family.Plant materialThe use of plants in the present study complies with international, national and/or institutional guidelines. All permissions to collect T. melanosporum fruitbodies in truffle orchards were obtained. The formal identification of biological material used in the study (T. melanosporum fruitbodies) was undertaken by F. Richard and E. Taschen. Voucher specimens of all collected fruitbodies have been deposited in the Centre d’Ecologie Fonctionnelle et Evolutive herbarium in Montpellier (France).Ethical approvalAll co-authors approve the ethical statement regarding the submitted manuscript.Consent to participateAll co-authors consent to participate to the research and agree with the content of the submitted manuscript. All authors reviewed and submitted manuscript. More

  • in

    Spatial distribution and interactions between mosquitoes (Diptera: Culicidae) and climatic factors in the Amazon, with emphasis on the tribe Mansoniini

    Changes in temperature and extreme environmental conditions can affect the dynamics of vector-borne pathogens. These include leishmaniasis, transmitted by phlebotomine sandflies, as well as mosquitoes that spread arboviruses like dengue, encephalitis, yellow fever, West Nile fever, and lymphatic filariasis19,20,21.The CCA analysis showed that maximum temperature significantly influenced the abundance of mosquito populations in the study area. In addition, the NMDS showed two different groupings that consisted of samples collected during the rainy and dry seasons. Accordingly, Refs.22,23 report that changes in temperature and relative humidity determine the abundance of mosquitoes, which can disappear entirely during the dry season. Moreover, Refs.22,24,25 note that certain species of mosquitoes increase proportionally with the regional rainfall regime. This is consistent with Ref.10, who find alternating patterns in tropical and temperate climates in some Brazilian regions.As shown by the geometric regression, there is a positive correlation between cumulative rainfall in the days before collection and the number of species found in the study period. Likewise, Ref.26 reported that under the conditions observed in the Serra do Mar State Park, climate variables directly influenced the abundance of Cq. chrysonotum and Cq. venezuelensis, favoring the occurrence of culicids during the more warm, wet, and rainy months.The current climate scenario and future projections about climate, environmental, demographic, and meteorological factors directly influence the distribution and abundance of mosquito vectors and/or diseases27,28,29,30. Environmental temperature alters mosquito population dynamics, thereby affecting the development of immature stages as well as reproduction31. While temperature has an important effect on population dynamics, rainfall and drought also affect the density and dispersal of mosquitoes in temperate and tropical regions32.To be sure, environmental changes other than climate can modify the behavior of vector insects and, subsequently, the mechanism of transmission of parasites20. Specifically, human impacts on the environment can result in drastically different disease transmission cycles in and around inhabited areas33.A previous study34 reported that changes in land use influence the mosquito communities with potential implications for the emergence of arboviruses. Another study35 noted that environmental changes negatively affect natural ecosystems with accelerated biodiversity loss. This is due to the modification and loss of natural habitat and unsustainable land use, which leads to the spread of pathogens and disease vectors.Hence, understanding the relationship between humans and the environment becomes increasingly critical, given the way in which climate changes can lead to alterations in the epidemiology of diseases such as dengue in areas considered free of the disease, as well as in endemic areas36.We found that the abundance and diversity of Mansoniini were directly influenced by the effect of the rainy season and other climatic factors. The rainfall regime has been shown to affect the development of immature forms12,37; explaining the greater frequency of these specimens in the warmer and wetter months38,39,40. According to Ref.41, stable ecosystems such as forests contain great species diversity. On the other hand, diversity tends to be reduced in biotic communities suffering from stress.Studies of insect populations in natural areas are important because they allow a direct analysis of how environmental factors influence phenomena such as the choice of breeding sites by females for oviposition, hematophagous behavior, and the distribution of species along a vegetation gradient12,26,42,43.Throughout the experimental period of the present study, we observed that Shannon light traps are an effective method for catching mosquitoes from the Mansoniini tribe. Interestingly, Ref.44 reported a species richness pattern strongly influenced by Coquillettidia fasciolata (Lynch Arribálzaga, 1891) on mosquito samples from different capture points by using CDC and Shannon light traps as sampling methods. In contrast to the results of Ref.44, where the highest population density of mosquitoes was captured with CDC traps, we observed that these traps were not effective at capturing specimens of Mansoniini in spite of being used in large numbers in the present study. Moreover, Ref.45 conducted another study on faunal diversity in an Atlantic Forest remnant of the state of Rio de Janeiro and observed the highest abundance of Cq. chrysonotum (Peryassú, 1922) and Cq. venezuelensis by using Shannon light traps, while the numbers of captures of Ma. titillans were very similar using CDC and Shannon traps.The results of this study indicate that the makeup of culicid fauna remains quite similar throughout the year, despite seasonal variations in abundance, though there was a lower variability of fauna in the dry season. Therefore, although the seasonality did not affect the temporal variation of the faunal composition in a generalized way, it was possible to detect a partial effect of the seasonality on fauna abundance.
    Reference46 report that the incidence peaks of mosquitoes in the warmer and wetter months, as well as mosquito populations remaining between tolerance limits for most of the year, indicate the sensitivity of some species to the local climate.The elevated abundance and diversity of species of Mansoniini in the study area were influenced by the favorable maintenance of breeding sites, including specific water accumulations with emerging vegetation that remain present throughout the year and the well-defined rainy season in the region. In addition, the representatives of Mansoniini, which prefer breeding sites containing macrophytes, made up nearly all of the species collected7.Besides providing a greater awareness of mosquito populations’ ecological and biological aspects, research carried out in wild areas also provides information on the relationship between species diversity and the area in which they are found. Considering that wild insects may become potential vectors of diseases, research in wild areas also provides helpful information for understanding relevant epidemiological aspects. These studies facilitate the identification, monitoring, and control of mosquito populations following environmental changes caused by direct human action, which can lead to major epidemics26.We observed considerable heterogeneity among Mansoniini fauna, and the months with the highest rainfall directly influence the structure of the communities and contribute to the increase in mosquito diversity and abundance, possibly due to variations in the availability of habitat for their immature forms. More

  • in

    Predicting potential global and future distributions of the African armyworm (Spodoptera exempta) using species distribution models

    Zeder, M. A. The domestication of animals. J. Anthropol. Res. 68, 161–190 (2012).Article 

    Google Scholar 
    Zohary, D. & Hopf, M. Domestication of Plants in the Old World: The Origin and Spread of Cultivated Plants in West Asia, Europe and the Nile Valley (Oxford University Press, 2000).
    Google Scholar 
    Epanchin-Niell, R., McAusland, C., Liebhold, A., Mwebaze, P. & Springborn, M. R. Biological invasions and international trade: Managing a moving target. Rev. Environ. Econom. Policy 15, 180–190 (2021).Article 

    Google Scholar 
    Gippet, J. M. & Bertelsmeier, C. Invasiveness is linked to greater commercial success in the global pet trade. Proc. Natl. Acad. Sci. 118, e2016337118 (2021).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Bertelsmeier, C. Globalization and the anthropogenic spread of invasive social insects. Curr. Opin. Insect Sci. 46, 16–23 (2021).PubMed 
    Article 

    Google Scholar 
    Charles, H. & Dukes, J. S. Biological Invasions 217–237 (Springer, 2008).
    Google Scholar 
    Bellard, C., Cassey, P. & Blackburn, T. M. Alien species as a driver of recent extinctions. Biol. Let. 12, 20150623 (2016).Article 

    Google Scholar 
    Bertolino, S. et al. Spatially explicit models as tools for implementing effective management strategies for invasive alien mammals. Mamm. Rev. 50, 187–199 (2020).Article 

    Google Scholar 
    Grimaldi, D., Engel, M. S., Engel, M. S. & Engel, M. S. Evolution of the Insects (Cambridge University Press, 2005).MATH 

    Google Scholar 
    Hill, M. P., Clusella-Trullas, S., Terblanche, J. S. & Richardson, D. M. Vol. 18, 883–891 (Springer, 2016).Sawicka, B. & Egbuna, C. Natural Remedies for Pest, Disease and Weed Control 1–16 (Elsevier, 2020).Book 

    Google Scholar 
    de la Vega, G. J. & Corley, J. C. Drosophila suzukii (Diptera: Drosophilidae) distribution modelling improves our understanding of pest range limits. Int. J. Pest Manag. 65, 217–227 (2019).Article 

    Google Scholar 
    Kriticos, D. J. et al. The potential distribution of invading Helicoverpa armigera in North America: Is it just a matter of time? PLoS ONE 10, e0119618 (2015).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Early, R., González-Moreno, P., Murphy, S. T. & Day, R. Forecasting the global extent of invasion of the cereal pest Spodoptera frugiperda, the fall armyworm. NeoBiota 40, 25–50 (2018).Article 

    Google Scholar 
    Day, R. et al. Fall armyworm: Impacts and implications for Africa. Outlooks Pest Manag. 28, 196–201 (2017).Article 

    Google Scholar 
    Rose, D. D. & Page, W. W. The African Armyworm Handbook 304 (Chatham, 2000).
    Google Scholar 
    De Groote, H. et al. Spread and impact of fall armyworm (Spodoptera frugiperda JE Smith) in maize production areas of Kenya. Agric. Ecosyst. Environ. 292, 106804 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Cheke, R. & Tucker, M. An evaluation of potential economic returns from the strategic control approach to the management of African armyworm Spodoptera exempta (Lepidoptera: Noctuidae) populations in eastern Africa. Crop Prot. 14, 91–103 (1995).Article 

    Google Scholar 
    Fox, K. Migrant Lepidoptera in New Zealand 1972–1973. N. Z. Entomol. 5, 268–271 (1973).Article 

    Google Scholar 
    Baker, G. An Outbreak of Spodoptera exempta (Walker) (Lepidoptera: Noctuidae) in the Highlands of Papua New Guinea (1978).Haggis, M. J. Distribution, Frequency of Attack and Seasonal Incidence of the African Armyworm Spodoptera exempta (Walk.) (Lep.: Noctuidae), with Particular Reference to Africa and Southwestern Arabia (Tropical Development and Research Institute, 1984).
    Google Scholar 
    Brown, E. Control of the African armyworm, Spodoptera exempta (Walk.)—An appreciation of the problem. East Afr. Agric. For. J. 35, 237–245 (1970).Article 

    Google Scholar 
    Rose, D. & Rainey, R. C. The significance of low-density populations of the African armyworm Spodoptera exempta (Walk.). Philos. Trans. R. Soc. Lond. B Biol. Sci. 287, 393–402 (1979).ADS 
    Article 

    Google Scholar 
    Tucker, M. & Pedgley, D. Rainfall and outbreaks of the African armyworm, Spodoptera exempta (Walker) (Lepidoptera: Noctuidae). Bull. Entomol. Res. 73, 195–199 (1983).Article 

    Google Scholar 
    Tucker, M. Forecasting the severity of armyworm seasons in East Africa from early season rainfall. Int. J. Trop. Insect Sci. 5, 51–55 (1984).Article 

    Google Scholar 
    Wilson, K. & Gatehouse, A. Seasonal and geographical variation in the migratory potential of outbreak populations of the African armyworm moth, Spodoptera exempta. J. Anim. Ecol. 62, 169–181 (1993).Article 

    Google Scholar 
    Odiyo, P. O. Development of the first outbreaks of the African armyworm, Spodoptera exempta (Walk.), between Kenya and Tanzania during the ‘off-season’ months of July to December. Int. J. Trop. Insect Sci. 1, 305–318 (1981).Article 

    Google Scholar 
    Haggis, M. Forecasting the severity of seasonal outbreaks of African armyworm, Spodoptera exempta (Lepidoptera: Noctuidae) in Kenya from the previous year’s rainfall. Bull. Entomol. Res. 86, 129–136 (1996).Article 

    Google Scholar 
    Harvey, A. & Mallya, G. Predicting the severity of Spodoptera exempta (Lepidoptera: Noctuidae) outbreak seasons in Tanzania. Bull. Entomol. Res. 85, 479–487 (1995).Article 

    Google Scholar 
    Holt, J., Mushobozi, W., Tucker, M. & Venn, J. Workshop on Research Priorities for Migrant Pests of Agriculture in Southern Africa, 151.Matthew Hill, T. C. M. Bloomberg (Online, 2017).Wilson, K. The Conversation (United Kingdom, 2017).Day, R. K. et al. WormBase: A data management and information system for forecasting Spodoptera exempta (Lepidoptera: Noctuidae) in eastern Africa. J. Econ. Entomol. 89, 1–10 (1996).Article 

    Google Scholar 
    Guisan, A. & Thuiller, W. Predicting species distribution: Offering more than simple habitat models. Ecol. Lett. 8, 993–1009 (2005).PubMed 
    Article 

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

    Google Scholar 
    Bosso, L. et al. The rise and fall of an alien: Why the successful colonizer Littorina saxatilis failed to invade the Mediterranean Sea. Biol. Invas. https://doi.org/10.1007/s10530-022-02838-y (2022).Article 

    Google Scholar 
    Sutherst, R. W. Pest species distribution modelling: Origins and lessons from history. Biol. Invas. 16, 239–256 (2014).Article 

    Google Scholar 
    Méndez-Vázquez, L. J., Lira-Noriega, A., Lasa-Covarrubias, R. & Cerdeira-Estrada, S. Delineation of site-specific management zones for pest control purposes: Exploring precision agriculture and species distribution modeling approaches. Comput. Electron. Agric. 167, 105101 (2019).Article 

    Google Scholar 
    Raffini, F. et al. From nucleotides to satellite imagery: Approaches to identify and manage the invasive pathogen Xylella fastidiosa and its insect vectors in Europe. Sustainability 12, 4508 (2020).CAS 
    Article 

    Google Scholar 
    Araújo, M. B. et al. Standards for distribution models in biodiversity assessments. Sci. Adv. 5, 4858 (2019).ADS 
    Article 

    Google Scholar 
    Hosmer, D. W. Jr., Lemeshow, S. & Sturdivant, R. X. Applied Logistic Regression Vol. 398 (Wiley, 2013).MATH 
    Book 

    Google Scholar 
    Landis, J. R. & Koch, G. G. The measurement of observer agreement for categorical data. Biometrics 33, 159–174 (1977).CAS 
    PubMed 
    MATH 
    Article 

    Google Scholar 
    Kalisa, W. et al. Assessment of climate impact on vegetation dynamics over East Africa from 1982 to 2015. Sci. Rep. 9, 1–20 (2019).ADS 
    CAS 
    Article 

    Google Scholar 
    Beck, H. E. et al. Present and future Köppen-Geiger climate classification maps at 1-km resolution. Sci. Data 5, 1–12 (2018).ADS 
    Article 

    Google Scholar 
    Mayaux, P., Bartholomé, E., Fritz, S. & Belward, A. A new land-cover map of Africa for the year 2000. J. Biogeogr. 31, 861–877 (2004).Article 

    Google Scholar 
    Marchant, R. et al. Drivers and trajectories of land cover change in East Africa: Human and environmental interactions from 6000 years ago to present. Earth Sci. Rev. 178, 322–378 (2018).ADS 
    Article 

    Google Scholar 
    Elith, J., Kearney, M. & Phillips, S. The art of modelling range-shifting species. Methods Ecol. Evol. 1, 330–342 (2010).Article 

    Google Scholar 
    Pemberton, C. E. Highlights in the history of entomology in Hawaii 1778–1963. Pac. Insects 6, 689–729 (1964).
    Google Scholar 
    Andow, D. A. Vegetational diversity and arthropod population response. Annu. Rev. Entomol. 36, 561–586 (1991).Article 

    Google Scholar 
    Andow, D. The extent of monoculture and its effects on insect pest populations with particular reference to wheat and cotton. Agr. Ecosyst. Environ. 9, 25–35 (1983).Article 

    Google Scholar 
    Oliveira, C., Auad, A., Mendes, S. & Frizzas, M. Crop losses and the economic impact of insect pests on Brazilian agriculture. Crop Prot. 56, 50–54 (2014).Article 

    Google Scholar 
    Furlong, M. J., Wright, D. J. & Dosdall, L. M. Diamondback moth ecology and management: Problems, progress, and prospects. Annu. Rev. Entomol. 58, 517–541 (2013).CAS 
    PubMed 
    Article 

    Google Scholar 
    Howse, M. W., Haywood, J. & Lester, P. J. Bioclimatic modelling identifies suitable habitat for the establishment of the invasive European paper wasp (Hymenoptera: Vespidae) across the southern hemisphere. Insects 11, 784 (2020).PubMed Central 
    Article 

    Google Scholar 
    Rose, D., Dewhurst, C., Page, W. & Fishpool, L. The role of migration in the life system of the African armyworm Spodoptera exempta. Int. J. Trop. Insect Sci. 8, 561–569 (1987).Article 

    Google Scholar 
    Dewhurst, C. F., Page, W. W. & Rose, D. J. The relationship between outbreaks, rainfall and low density populations of the African armyworm, Spodoptera exempta, Kenya. Entomol. Exp. et Appl. 98, 285–294 (2001).Article 

    Google Scholar 
    Aguilon, D. J. & Velasco, L. R. Effects of larval rearing temperature and host plant condition on the development, survival, and coloration of African armyworm, Spodoptera exempta Walker (Lepidoptera: Noctuidae). J. Environ. Sci. Manag. 18, 54 (2015).Article 

    Google Scholar 
    David, W. & Ellaby, S. The viability of the eggs of the African army-worm, Spodoptera exempta in laboratory cultures. Entomol. Exp. Appl. 18, 269–280 (1975).Article 

    Google Scholar 
    He, L., Zhao, S., Ali, A., Ge, S. & Wu, K. Ambient humidity affects development, survival, and reproduction of the invasive fall armyworm, Spodoptera frugiperda (Lepidoptera: Noctuidae), China. J. Econ. Entomol. 114, 1145–1158 (2021).PubMed 
    Article 

    Google Scholar 
    Janssen, J. Effects of the mineral composition and water content of intact plants on the fitness of the African armyworm. Oecologia 95, 401–409 (1993).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Shahzad, M. S. et al. Modelling population dynamics of army worm (Spodoptera litura F.) (Lepidoptera: Noctuiidae) in relation to meteorological factors in Multan, Punjab, Pakistan. Int. J. Agron. Agric. Res. 5, 39–45 (2014).
    Google Scholar 
    Garcia, A. G., Ferreira, C. P., Godoy, W. A. & Meagher, R. L. A computational model to predict the population dynamics of Spodoptera frugiperda. J. Pest. Sci. 92, 429–441 (2019).Article 

    Google Scholar 
    Hickling, R., Roy, D. B., Hill, J. K., Fox, R. & Thomas, C. D. The distributions of a wide range of taxonomic groups are expanding polewards. Glob. Change Biol. 12, 450–455 (2006).ADS 
    Article 

    Google Scholar 
    Vanhanen, H., Veteli, T. O., Paivinen, S., Kellomaki, S. & Niemela, P. Climate change and range shifts in two insect defoliators: Gypsy moth and nun moth-a model study. Silva Fennica 41, 621 (2007).Article 

    Google Scholar 
    Falk, W. & Hempelmann, N. Species favourability shift in Europe due to climate change: A case study for Fagus sylvatica L. and Picea abies (L.) Karst. based on an ensemble of climate models. J. Climatol. 2013, 1–18 (2013).Article 

    Google Scholar 
    Arora, R. & Dhawan, A. Climate Change and Insect Pest Management. Integrated Pest Management 44–60 (Scientific Publisher, 2013).
    Google Scholar 
    Andrew, N. R. & Hill, S. J. Effect of climate change on insect pest management. In Environmental Pest Management: Challenges for Agronomists, Ecologists, Economists and Policymakers, 197 (2017).De Boer, J. G. & Harvey, J. A. Range-expansion in processionary moths and biological control. Insects 11, 267 (2020).PubMed Central 
    Article 

    Google Scholar 
    Bras, A. et al. A complex invasion story underlies the fast spread of the invasive box tree moth (Cydalima perspectalis) across Europe. J. Pest. Sci. 92, 1187–1202 (2019).Article 

    Google Scholar 
    Araújo, M. B. et al. Heat freezes niche evolution. Ecol. Lett. 16, 1206–1219 (2013).PubMed 
    Article 

    Google Scholar 
    Barford, E. Crop pests advancing with global warming. Nature 10, 13644 (2013).
    Google Scholar 
    Bebber, D. P., Ramotowski, M. A. & Gurr, S. J. Crop pests and pathogens move polewards in a warming world. Nat. Clim. Change 3, 985–988 (2013).ADS 
    Article 

    Google Scholar 
    Rubenstein, D. I. The greenhouse effect and changes in animal behavior: Effects on social structure and life-history strategies. In Global Warming and Biological Diversity, 180–192 (1992).Karuppaiah, V. & Sujayanad, G. Impact of climate change on population dynamics of insect pests. World J. Agric. Sci. 8, 240–246 (2012).
    Google Scholar 
    Jakhar, B. et al. Influence of climate change on Helicoverpa armigera (Hubner) in pigeonpea. J. Agric. Ecol. 2, 25–31 (2016).
    Google Scholar 
    Akbar, S. M., Pavani, T., Nagaraja, T. & Sharma, H. Influence of CO 2 and temperature on metabolism and development of Helicoverpa armigera (Noctuidae: Lepidoptera). Environ. Entomol. 45, 229–236 (2016).CAS 
    PubMed 
    Article 

    Google Scholar 
    Magandana, T. P., Hassen, A. & Tesfamariam, E. H. Seasonal herbaceous structure and biomass production response to rainfall reduction and resting period in the semi-arid grassland area of South Africa. Agronomy 10, 1807 (2020).CAS 
    Article 

    Google Scholar 
    Gherardi, L. A. & Sala, O. E. Enhanced precipitation variability decreases grass-and increases shrub-productivity. Proc. Natl. Acad. Sci. 112, 12735–12740 (2015).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Scheiter, S. & Higgins, S. I. Impacts of climate change on the vegetation of Africa: An adaptive dynamic vegetation modelling approach. Glob. Change Biol. 15, 2224–2246 (2009).ADS 
    Article 

    Google Scholar 
    Hernandez, P. A., Graham, C. H., Master, L. L. & Albert, D. L. The effect of sample size and species characteristics on performance of different species distribution modeling methods. Ecography 29, 773–785 (2006).Article 

    Google Scholar 
    Jiménez-Valverde, A., Lobo, J. & Hortal, J. The effect of prevalence and its interaction with sample size on the reliability of species distribution models. Community Ecol. 10, 196–205 (2009).Article 

    Google Scholar 
    Renault, D., Laparie, M., McCauley, S. J. & Bonte, D. Environmental adaptations, ecological filtering, and dispersal central to insect invasions. Annu. Rev. Entomol. 63, 345–368 (2018).CAS 
    PubMed 
    Article 

    Google Scholar 
    Ellis, S. New pest response guidelines: Spodoptera. USDA/APHIS/PPQ/PDMP (2004).Waage, J. & Mumford, J. D. Agricultural biosecurity. Philos. Trans. R. Soc. B Biol. Sci. 363, 863–876 (2008).CAS 
    Article 

    Google Scholar 
    Anand, M. A systems approach to agricultural biosecurity. Health Secur. 16, 58–68 (2018).PubMed 
    Article 

    Google Scholar 
    MacLeod, A., Pautasso, M., Jeger, M. J. & Haines-Young, R. Evolution of the international regulation of plant pests and challenges for future plant health. Food Secur. 2, 49–70 (2010).Article 

    Google Scholar 
    Jiménez-Valverde, A. et al. Use of niche models in invasive species risk assessments. Biol. Invas. 13, 2785–2797 (2011).Article 

    Google Scholar 
    Oluwole, F. A., Sambo, J. M. & Sikhalazo, D. Long-term effects of different burning frequencies on the dry savannah grassland in South Africa. Afr. J. Agric. Res. 3, 147–153 (2008).
    Google Scholar 
    Kalleshwaraswamy, C. et al. First Report of the Fall Armyworm, Spodoptera frugiperda (JE Smith) (Lepidoptera: Noctuidae), an Alien Invasive Pest on Maize in India (2018).Bentivenha, J., Baldin, E., Hunt, T., Paula-Moraes, S. & Blankenship, E. Intraguild competition of three noctuid maize pests. Environ. Entomol. 45, 999–1008 (2016).CAS 
    PubMed 
    Article 

    Google Scholar 
    Chapman, J. W. et al. Fitness consequences of cannibalism in the fall armyworm, Spodoptera frugiperda. Behav. Ecol. 10, 298–303 (1999).Article 

    Google Scholar 
    Divya, J., Kalleshwaraswamy, C., Mallikarjuna, H. & Deshmukh, S. Does recently invaded fall armyworm, Spodoptera frugiperda displace native lepidopteran pests of maize in India? Curr. Sci. 120, 1358 (2021).Article 

    Google Scholar 
    Hailu, G. et al. Could fall armyworm, Spodoptera frugiperda (JE Smith) invasion in Africa contribute to the displacement of cereal stemborers in maize and sorghum cropping systems. Int. J. Trop. Insect Sci. 41, 1753–1762 (2021).Article 

    Google Scholar 
    Srivastava, V., Lafond, V. & Griess, V. C. Species distribution models (SDM): Applications, benefits and challenges in invasive species management. CAB Rev. 14, 1–13 (2019).Article 

    Google Scholar 
    Fick, S. E. & Hijmans, R. J. WorldClim 2: New 1-km spatial resolution climate surfaces for global land areas. Int. J. Climatol. 37, 4302–4315 (2017).Article 

    Google Scholar 
    Eyring, V. et al. Overview of the coupled model intercomparison project phase 6 (CMIP6) experimental design and organization. Geosci. Model Dev. 9, 1937–1958 (2016).ADS 
    Article 

    Google Scholar 
    Wu, T. et al. The Beijing Climate Center climate system model (BCC-CSM): The main progress from CMIP5 to CMIP6. Geosci. Model Dev. 12, 1573–1600 (2019).ADS 
    Article 

    Google Scholar 
    O’Neill, B. C. et al. The roads ahead: Narratives for shared socioeconomic pathways describing world futures in the 21st century. Glob. Environ. Change 42, 169–180 (2017).Article 

    Google Scholar 
    Petitpierre, B., Broennimann, O., Kueffer, C., Daehler, C. & Guisan, A. Selecting predictors to maximize the transferability of species distribution models: Lessons from cross-continental plant invasions. Glob. Ecol. Biogeogr. 26, 275–287 (2017).Article 

    Google Scholar 
    Cano, J. et al. Modelling the spatial distribution of aquatic insects (Order Hemiptera) potentially involved in the transmission of Mycobacterium ulcerans in Africa. Parasit. Vectors 11, 1–16 (2018).Article 

    Google Scholar 
    Gómez-Undiano, I. Modelos y patrones de distribución geográfica de especies de Culicidae (Culex pipiens, Mansonia africana y Mansonia uniformis) vectores de filariasis linfática en ámbitos urbanos y periurbanos del África subsahariana. Máster en Zoología thesis, Universidad Complutense de Madrid (2018).R: A Language and Environment for Statistical Computing. (R Foundation for Statistical Computing, 2020).Phillips, S. J. et al. Sample selection bias and presence-only distribution models: Implications for background and pseudo-absence data. Ecol. Appl. 19, 181–197 (2009).PubMed 
    Article 

    Google Scholar 
    Elith, J. et al. Novel methods improve prediction of species’ distributions from occurrence data. Ecography 29, 129–151 (2006).Article 

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

    Google Scholar 
    Thuiller, W., Lafourcade, B., Engler, R. & Araújo, M. B. BIOMOD—A platform for ensemble forecasting of species distributions. Ecography 32, 369–373 (2009).Article 

    Google Scholar 
    Thuiller, W. et al. Package ‘biomod2’. Species Distribution Modeling Within an Ensemble Forecasting Framework (2016).Acevedo, P., Jiménez-Valverde, A., Lobo, J. M. & Real, R. Delimiting the geographical background in species distribution modelling. J. Biogeogr. 39, 1383–1390 (2012).Article 

    Google Scholar 
    VanDerWal, J., Shoo, L. P., Graham, C. & Williams, S. E. Selecting pseudo-absence data for presence-only distribution modeling: How far should you stray from what you know? Ecol. Model. 220, 589–594 (2009).Article 

    Google Scholar 
    Hijmans, R., Phillips, S., Leathwick, J. & Elith, J. (2012).Araújo, M. B. & New, M. Ensemble forecasting of species distributions. Trends Ecol. Evol. 22, 42–47 (2007).PubMed 
    Article 

    Google Scholar 
    Gama, M., Crespo, D., Dolbeth, M. & Anastácio, P. M. Ensemble forecasting of Corbicula fluminea worldwide distribution: Projections of the impact of climate change. Aquat. Conserv. Mar. Freshw. Ecosyst. 27, 675–684 (2017).Article 

    Google Scholar 
    Liu, C., White, M., Newell, G. & Griffioen, P. Species distribution modelling for conservation planning in Victoria, Australia. Ecol. Model. 249, 68–74 (2013).Article 

    Google Scholar  More