More stories

  • in

    Water motion and pH jointly impact the availability of dissolved inorganic carbon to macroalgae

    Duggins, D. O., Simenstad, C. A. & Estes, J. A. Magnification of secondary producition by kelp detritus in coastal marine ecosystems. Science 1979(245), 170–173 (1989).Article 
    ADS 

    Google Scholar 
    Hill, R. et al. Can macroalgae contribute to blue carbon? An Australian perspective. Limnol. Oceanogr. 60, 1689–1706 (2015).Article 
    ADS 

    Google Scholar 
    Mann, K. H. Seaweeds: Their productivity and strategy for growth. Science 1979(182), 975–981 (1973).Article 
    ADS 

    Google Scholar 
    Steneck, R. S. et al. Kelp forest ecosystems: Biodiversity, stability, resilience and future. Environ. Conserv. 29, 436–459 (2002).Article 

    Google Scholar 
    Giordano, M., Beardall, J. & Raven, J. A. CO2 concentrating mechanisms in algae: Mechanisms, environmental modulation, and evolution. Annu. Rev. Plant Biol. 56, 99–131 (2005).Article 
    CAS 

    Google Scholar 
    Raven, J. A. & Beardall, J. The ins and outs of CO2. J. Exp. Bot. 67, 1–13 (2016).Article 
    CAS 

    Google Scholar 
    Raven, J. A. et al. Seaweeds in cold seas: Evolution and carbon acquisition. Ann. Bot. 90, 525–536. https://doi.org/10.1093/aob/mcf171 (2002).Article 
    CAS 

    Google Scholar 
    Raven, J. et al. Ocean Acidification due to Increasing Atmospheric Carbon Dioxide 1–68 (The Royal Society, 2005).
    Google Scholar 
    Kübler, J. E. & Dudgeon, S. R. Predicting effects of ocean acidification and warming on algae lacking carbon concentrating mechanisms. PLoS ONE 10, 1–19 (2015).Article 

    Google Scholar 
    Fernández, P. A., Hurd, C. L. & Roleda, M. Y. Bicarbonate uptake via an anion excange protein is the main mechanism of inorganic carbon acquisition by the giant kelp Macrocystis pyrifera (Laminariales, Phaeophyceae) under variable pH1. J. Phycol. 50, 1–11 (2014).Article 

    Google Scholar 
    Raven, J. A. et al. Mechanistic interpretation of carbon isotope discrimination by marine macroalgae and seagrasses. Funct. Plant Biol. 29, 355 (2002).Article 
    CAS 

    Google Scholar 
    Raven, J. A., Cockell, C. S. & De La Rocha, C. L. The evolution of inorganic carbon concentrating mechanisms in photosynthesis. Philos. Trans. R. Soc. B 363, 2641–2650 (2008).Article 
    CAS 

    Google Scholar 
    Bidwell, R. G. S. S. & McLachlan, J. Carbon nutrition of seaweeds: Photosynthesis, photorespiration and respiration. J. Exp. Mar. Biol. Ecol. 86, 15–46 (1985).Article 
    CAS 

    Google Scholar 
    Hurd, C. L. Water motion, marine macroalgal physiology and production. J. Phycol. 36, 453–472. https://doi.org/10.1046/j.1529-8817.2000.99139.x (2000).Article 
    CAS 

    Google Scholar 
    Hurd, C. L., Stevens, C. L., Laval, B. E., Lawrence, G. A. & Harrison, P. J. Visualization of seawater flow around morphologically distinct forms of the giant kelp Macrocystis integrifolia from wave-sheltered and exposed sites. Limnol. Oceanogr. 42, 156–163. https://doi.org/10.4319/lo.1997.42.1.0156 (1997).Article 
    ADS 

    Google Scholar 
    Smith, F. A. A. & Walker, N. A. A. Photosynthesis by aquatic plants: Effects of unstirred layers in relation to assimilation of CO2 and HCO3- to carbon isotope discrimination. N. Phytol. 86, 245–259 (1980).Article 
    CAS 

    Google Scholar 
    Wheeler, W. N. Effect of boundary layer transport on the fixation of carbon by the giant kelp Macrocystis pyrifera. Mar. Biol. 56, 103–110 (1980).Article 
    ADS 
    CAS 

    Google Scholar 
    Hurd, C. L., Lenton, A., Tilbrook, B. & Boyd, P. W. Current understanding and challenges for oceans in a higher-CO2 world. Nat. Clim. Chang. 8, 686–694 (2018).Article 
    ADS 
    CAS 

    Google Scholar 
    Stocker, T. F. et al. Technical Summary. Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change 33–115 (2013).Hepburn, C. D. et al. Diversity of carbon use strategies in a kelp forest community: Implications for a high CO2 ocean. Glob. Chang. Biol. 17, 2488–2497 (2011).Article 
    ADS 

    Google Scholar 
    Beer, S. & Koch, E. Photosynthesis of marine macroalgae and seagrasses in globally changing CO2 environments. Mar. Ecol. Prog. Ser. 141, 199–204 (1996).Article 
    ADS 

    Google Scholar 
    Ihnken, S., Roberts, S. & Beardall, J. Differential responses of growth and photosynthesis in the marine diatom Chaetoceros muelleri to CO2 and light availability. Phycologia 50, 182–193 (2011).Article 
    CAS 

    Google Scholar 
    Gerard, V. A. In situ water motion and nutrient uptake by the giant kelp Macrocystis pyrifera. Mar. Biol. 69, 51–54 (1982).Article 

    Google Scholar 
    Hepburn, C. D., Holborow, J. D., Wing, S. R., Frew, R. D. & Hurd, C. L. Exposure to waves enhances the growth rate and nitrogen status of the giant kelp Macrocystis pyrifera. Mar. Ecol. Prog. Ser. 339, 99–108 (2007).Article 
    ADS 
    CAS 

    Google Scholar 
    Hurd, C. L. Shaken and stirred: The fundamental role of water motion in resource acquisition and seaweed productivity. Persp. Phycol. 4, 73–81 (2017).ADS 

    Google Scholar 
    Sültemeyer, D. F., Miller, A. G., Espie, G. S., Fock, H. P. & Canvin, D. T. Active CO2 transport by the green alga Chlamydomonas reinhardtii. Plant Physiol. 89, 1213–1219 (1989).Article 

    Google Scholar 
    Koch, M., Bowes, G., Ross, C. & Zhang, X. H. Climate change and ocean acidification effects on seagrasses and marine macroalgae. Glob. Chang. Biol. 19, 103–132 (2013).Article 
    ADS 

    Google Scholar 
    Britton, D., Cornwall, C. E., Revill, A. T., Hurd, C. L. C. L. & Johnson, C. R. Ocean acidification reverses the positive effects of seawater pH fluctuations on growth and photosynthesis of the habitat-forming kelp Ecklonia radiata. Sci. Rep. 6, 1–10 (2016).Article 

    Google Scholar 
    Cornwall, C. E. et al. Carbon-use strategies in macroalgae: Differential responses to lowered ph and implications for ocean acidification. J. Phycol. 48, 137–144 (2012).Article 
    CAS 

    Google Scholar 
    Kram, S. L. et al. Variable responses of temperate calcified and fleshy macroalgae to elevated pCO2 and warming. ICES J. Mar. Sci. 73, 693–703 (2016).Article 

    Google Scholar 
    Kübler, J. E., Johnston, A. M. & Raven, J. A. The effects of reduced and elevated CO2 and O2 on the seaweed Lomentaria articulata. Plant Cell Environ. 22, 1303–1310 (1999).Article 

    Google Scholar 
    van der Loos, L. M. et al. Responses of macroalgae to CO2 enrichment cannot be inferred solely from their inorganic carbon uptake strategy. Ecol. Evol. 9, 125–140 (2019).Article 

    Google Scholar 
    Cornwall, C. E. & Hurd, C. L. Variability in the benefits of ocean acidification to photosynthetic rates of macroalgae without CO2-concentrating mechanisms. Mar. Freshw. Res. 71, 275–280 (2019).Article 

    Google Scholar 
    Cornwall, C. E., Revill, A. T. & Hurd, C. L. High prevalence of diffusive uptake of CO2 by macroalgae in a temperate subtidal ecosystem. Photosynth. Res. 124, 181–190 (2015).
    Article 
    CAS 

    Google Scholar 
    Lovelock, C. E., Reef, R., Raven, J. A. & Pandolfi, J. M. Regional variation in δ13C of coral reef macroalgae. Limnol. Oceanogr. 65, 2291–2302 (2020).Article 
    ADS 
    CAS 

    Google Scholar 
    Fischer, G. & Wiencke, C. Stable carbon isotope composition, depth distribution and fate of macroalgae from the Antarctic Peninsula region. Polar. Biol. 12, 341–348 (1992).Article 

    Google Scholar 
    Stephens, T. A. & Hepburn, C. D. Mass-transfer gradients across kelp beds influence Macrocystis pyrifera growth over small spatial scales. Mar. Ecol. Prog. Ser. 515, 97–109 (2014).Article 
    ADS 

    Google Scholar 
    Kregting, L. T., Hepburn, C. D. & Savidge, G. Seasonal differences in the effects of oscillatory and uni-directional flow on the growth and nitrate-uptake rates of juvenile Laminaria digitata (Phaeophyceae). J. Phycol. 51, 1116–1126 (2015).Article 
    CAS 

    Google Scholar 
    Parker, H. S. Influence of relative water motion on the growth, ammonium uptake and carbon and nitrogen composition of Ulva lactuca (Chlorophyta). Mar. Biol. 63, 309–318 (1981).Article 
    CAS 

    Google Scholar 
    Bergstrom, E. et al. Inorganic carbon uptake strategies in coralline algae: Plasticity across evolutionary lineages under ocean acidification and warming. Mar. Environ. Res. 161, 105107 (2020).Article 
    CAS 

    Google Scholar 
    Maberly, S. C., Raven, J. A. & Johnston, A. M. Discrimination between C-12 and C-13 by marine plants. Oecologia 91, 481–492 (1992).Article 
    ADS 
    CAS 

    Google Scholar 
    Gattuso, J. P. et al. Package ‘Seacarb ’. Preprint at http://cran.r-project.org/package=seacarb (2015).Raven, J. A., Beardall, J. & Giordano, M. Energy costs of carbon dioxide concentrating mechanisms in aquatic organisms. Photosynth. Res. 121, 111–124 (2014).Article 
    CAS 

    Google Scholar 
    Raven, J. A., Walker, D. I., Johnston, A. M., Handley, L. L. & Kübler, J. E. Implications of 13C natural abundance measurements for photosynthetic performance by marine macrophytes in their natural environment. Mar. Ecol. Prog. Ser. 123, 193–205 (1995).Article 
    ADS 

    Google Scholar 
    Raven, J. A. Inorganic carbon acquisition by marine autotrophs. Adv. Bot. Res. 27, 85–209 (1997).Article 
    CAS 

    Google Scholar 
    Fernández, P. A., Roleda, M. Y. & Hurd, C. L. Effects of ocean acidification on the photosynthetic performance, carbonic anhydrase activity and growth of the giant kelp Macrocystis pyrifera. Photosynth. Res. 124, 293–304 (2015).Article 

    Google Scholar 
    Bailly, J. & Coleman, J. R. Effect of CO(2) concentration on protein biosynthesis and carbonic anhydrase expression in Chlamydomonas reinhardtii. Plant Physiol. 87, 833–840 (1988).Article 
    CAS 

    Google Scholar 
    Dionisio-Sese, M. L., Fukuzawa, H. & Miyachi, S. Light-induced carbonic anhydrase expression in Chlamydomonas reinhardtii. Plant Physiol. 94, 1103–1110 (1990).Article 
    CAS 

    Google Scholar 
    Pollock, S. V., Colombo, S. L., Prout, D. L., Godfrey, A. C. & Moroney, J. V. Rubisco activase is required for optimal photosynthesis in the green alga Chlamydomonas reinhardtii in a low-CO2 atmosphere. Plant Physiol. 133, 1854–1861 (2003).Article 
    CAS 

    Google Scholar 
    Carlberg, S., Axelsson, L., Larsson, C., Ryberg, H. & Uusitalo, J. Inducible CO2 concentrating mechanisms in green seaweeds I. Taxonomical and physiological aspects. In Current Research in Photosynthesis (ed. Baltscheffsky, M.) (Springer, 1990). https://doi.org/10.1007/978-94-009-0511-5_749.Chapter 

    Google Scholar 
    Wheeler, W. N. Effect of boundary-layer transport on the fixation of carbon by the giant-kelp Macrocystis pyrifera. Mar. Biol. 56, 103–110 (1980).Article 
    ADS 
    CAS 

    Google Scholar 
    Johnston, A. M. & Raven, J. A. Effects of culture in high CO2 on the photosynthetic physiology of Fucus serratus. Br. J. Phycol. 25, 75–82 (1990).Article 

    Google Scholar 
    Connell, S. D., Kroeker, K. J., Fabricius, K. E., Kline, D. I. & Russell, B. D. The other ocean acidification problem: CO2 as a resource among competitors for ecosystem dominance. Philos. Trans. R. Soc. Lond. 368, 20120442 (2013).Article 

    Google Scholar 
    Porter, E. T., Sanford, L. P. & Suttles, S. E. Gypsum dissolution is not a universal integrator of water motion. Limnol. Oceanogr. 45, 145–158 (2000).Article 
    ADS 

    Google Scholar 
    Gerard, V. A. & Mann, K. H. Growth and production of Laminaria longicruris (Phaeophyta) populations exposed to different intensities of water movement. J. Phycol. 15, 33–41 (1979).Article 

    Google Scholar 
    Bivand, R., Keitt, T. & Rowlingson, B. Package ‘rgdal’. R Package https://doi.org/10.1353/lib.0.0050 (2016).Article 

    Google Scholar 
    LINZ. LINZ Data Service. https://data.linz.govt.nz/layer/50258-nz-coastlines-topo-150k/history/ Accessed July 2021 (2021).Kirk, J. T. Characteristics of the light field in highly turbid waters: A Monte Carlo study. Limnol. Oceanogr. 39, 702–706 (1994).Article 
    ADS 

    Google Scholar 
    Strickland, J. D. H. & Parsons, T. R. A Practical Handbook of Seawater Analysis (Fisheries Research Board of Canada, 1968).
    Google Scholar 
    Kohler, K. E. & Gill, S. M. Coral Point Count with Excel extensions (CPCe): A visual basic program for the determination of coral and substrate coverage using random point count methodology. Comput. Geosci. 32, 1259–1269 (2006).Article 
    ADS 

    Google Scholar 
    Axelsson, L., Mercado, J. & Figueroa, F. Utilization of HCO3− at high ph by the brown macroalga laminaria saccharina. Eur. J. Phycol. 35, 53–59 (2000).Article 

    Google Scholar 
    R Core Team. R: A language and environment for statistical computing. Preprint at (2017).Bates, D., Maechler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1–48 (2015).Article 

    Google Scholar 
    Hothorn, T., Bretz, F. & Westfall, P. Simultaneous inference in general parametric models. Biom. J. 50, 346–363 (2008).Article 
    MathSciNet 
    MATH 

    Google Scholar  More

  • in

    Neolithic dental calculi provide evidence for environmental proxies and consumption of wild edible fruits and herbs in central Apennines

    Asevedo, L. et al. Palynological analysis of dental calculus from Pleistocene proboscideans of southern Brazil: a new approach for paleodiet and paleoenvironmental reconstructions. Palaeogeogr. Palaeoclimatol. Palaeoecol. 540, 109523 (2020).Article 

    Google Scholar 
    Cristiani, E. et al. Wild cereal grain consumption among Early Holocene foragers of the Balkans predates the arrival of agriculture. Elife 10, e72976 (2021).Article 
    CAS 

    Google Scholar 
    Nava, A. et al. Multipronged dental analyses reveal dietary differences in last foragers and first farmers at Grotta Continenza, central Italy (15,500–7000 BP). Sci. Rep. 11, 1–14 (2021).Article 

    Google Scholar 
    Ottoni, C. et al. Tracking the transition to agriculture in Southern Europe through ancient DNA analysis of dental calculus. Proc. Natl. Acad. Sci. USA 118, e2102116118 (2021).Article 
    CAS 

    Google Scholar 
    Cammidge, T. S., Kooyman, B. & Theodor, J. M. Diet reconstructions for end-Pleistocene Mammut americanum and Mammuthus based on comparative analysis of mesowear, microwear, and dental calculus in modern Loxodonta africana. Palaeogeogr. Palaeoclimatol. Palaeoecol. 538, 109403 (2020).Article 

    Google Scholar 
    de Oliveira, K. et al. From oral pathology to feeding ecology: the first dental calculus paleodiet study of a South American native megamammal. J. S. Am. Earth Sci. 109, 103281 (2021).Article 

    Google Scholar 
    Mothé, D. et al. The micro from mega: dental calculus description and the first record of fossilized oral bacteria from an extinct proboscidean. Int. J. Paleopathol. 33, 55–60 (2021).Article 

    Google Scholar 
    Eglinton, G. & Logan, G. A. Molecular preservation. Philosophical Transactions of the Royal Society of London. Philos. Trans. R. Soc. Lond., B, Biol. Sci. 333, 315–328 (1991).CAS 

    Google Scholar 
    Romanowski, G., Lorenz, M. G. & Wackernagel, W. Adsorption of plasmid DNA to mineral surfaces and protection against Dnase I. Appl. Environ. Microbiol. 57, 1057–1061 (1991).Article 
    CAS 

    Google Scholar 
    Milanesi, C. et al. Ultrastructural study of archaeological Vitis vinifera L. seeds using rapid-freeze fixation and substitution. Tissue Cell 41, 443–447 (2009).Article 
    CAS 

    Google Scholar 
    Power, R. C., Salazar-García, D. C., Wittig, R. M., Freiberg, M., & Henry, A. G. Dental calculus evidence of Taï Forest chimpanzee plant consumption and life history transitions. Sci. Rep. 5, 15161 (2015).Goude, G. et al. A multidisciplinary approach to Neolithic life reconstruction. J. Archaeol. Method Theory 26, 537–560 (2019).Article 

    Google Scholar 
    Farrer, A. G. et al. Effectiveness of decontamination protocols when analyzing ancient DNA preserved in dental calculus. Sci. Rep. 11, 1–14 (2021).Article 

    Google Scholar 
    Weyrich, L. S., Dobney, K. & Cooper, A. Ancient DNA analysis of dental calculus. J. Hum. Evol. 79, 119–124 (2015).Article 

    Google Scholar 
    Ozga, A. T. et al. Successful enrichment and recovery of whole mitochondrial genomes from ancient human dental calculus. Am. J. Phys. Anthropol. 160, 220–228 (2016).Article 

    Google Scholar 
    Mann, A. E. et al. Do I have something in my teeth? The trouble with genetic analyses of diet from archaeological dental calculus. Quat. Int. https://doi.org/10.1016/j.quaint.2020.11.019 (2020).Wright, S. L., Dobney, K. & Weyrich, L. S. Advancing and refining archaeological dental calculus research using multiomic frameworks. Sci. Technol. Archaeol. Res. 7, 13–30 (2021).
    Google Scholar 
    Sawafuji, R., Saso, A., Suda, W., Hattori, M. & Ueda, S. Ancient DNA analysis of food remains in human dental calculus from the Edo period, Japan. PLoS One 15, e0226654 (2020).Article 
    CAS 

    Google Scholar 
    Weyrich, L. S. et al. Neanderthal behaviour, diet, and disease inferred from ancient DNA in dental calculus. Nature 544, 357–361 (2017).Article 
    CAS 

    Google Scholar 
    Ottoni, C. et al. Metagenomic analysis of dental calculus in ancient Egyptian baboons. Sci. Rep. 9, 1–10 (2019).Article 

    Google Scholar 
    Hollingsworth, P. M., Graham, S. W. & Little, D. P. Choosing and using a Plant DNA barcode. PLoS One 6, 1–13 (2011).Article 

    Google Scholar 
    Gismondi, A., Fanali, F., Labarga, J. M. M., Caiola, M. G. & Canini, A. Crocus sativus L. genomics and different DNA barcode applications. Plant Syst. Evol. 299, 1859–1863 (2013).Article 
    CAS 

    Google Scholar 
    ICSN. The international code for starch nomenclature, accessed 15 September 2021; http://fossilfarm.org/ICSN/Code.html (2011).Gismondi, A. et al. Starch granules: a data collection of 40 food species. Plant Biosyst. 153, 273–279 (2019).Article 

    Google Scholar 
    Henry, A. G., Brooks, A. S. & Piperno, D. R. Plant foods and the dietary ecology of Neanderthals and early modern humans. J. Hum. Evol. 69, 44–54 (2014).Article 

    Google Scholar 
    PalDat. A palynological database (2000 onwards), accessed 19 January 2022; https://www.paldat.org/ (2019).Berglund, B. E. & Ralska-Jasiewiczowa, M. Pollen analysis and pollen diagrams. In Handbook of Holocene Palaeoecology and Palaeohydrology (ed. Berglund, B. E.) 455–484 (Wiley, 1986).Faegri, K. & Iversen, J. Textbook of Pollen analysis, 4th edn (eds Faegri, K. et al.) (John Wiley and Sons-Chichester, 1989).Grímsson, F. et al. Fagaceae pollen from the early Cenozoic of West Greenland: revisiting Engler’s and Chaney’s Arcto-Tertiary hypotheses. Plant Syst. Evol. 301, 809–832 (2015).Article 

    Google Scholar 
    Denk, T. & Tekleva, M. V. Pollen morphology and ultrastructure of Quercus with focus on Group Ilex (= Quercus Subgenus Heterobalanus (Oerst.) Menitsky): Implications for oak systematics and evolution. Grana 53, 255–282 (2014).Article 

    Google Scholar 
    Grímsson, F. & Zetter, R. Combined LM and SEM study of the middle Miocene (Sarmatian) palynofora from the Lavanttal Basin, Austria: Part II. Pinophyta (Cupressaceae, Pinaceae and Sciadopityaceae). Grana 50, 262–310 (2011).Article 

    Google Scholar 
    Mohanty, R. P., Buchheim, M. A., Portman, R. & Levetin, E. Molecular and ultrastructural detection of plastids in Juniperus (Cupressaceae) pollen. Phytologia 98, 298–310 (2016).
    Google Scholar 
    Martin, A. C. & Harvey, W. J. The Global Pollen Project: a new tool for pollen identifcation and the dissemination of physical reference collections. Methods Ecol. Evol. 8, 892–897 (2017).Article 

    Google Scholar 
    Maciejewska-Rutkowska, I., Bocianowski, J. & Wrońska-Pilarek, D. Pollen morphology and variability of Polish native species from genus Salix L. PLoS One 16, e0243993 (2021).Article 
    CAS 

    Google Scholar 
    Abreu, I., Costa, I., Oliveira, M., Cunha, M. & De Castro, R. Ultrastructure and germination of Vitis vinifera cv. Loureiro pollen. Protoplasma 228, 131–135 (2006).Article 
    CAS 

    Google Scholar 
    Nagels, A. et al. Palynological diversity and major evolutionary trends in Cyperaceae. Plant Syst. Evol. 277, 117 (2009).Article 

    Google Scholar 
    El Ghazali, G. E. Pollen morphological studies in Amaranthaceae s. lat. (incl. Chenopodiaceae) and their taxonomic significance: a review. Grana 61, 1–7 (2022).Article 

    Google Scholar 
    Petraco, N., & Kubic, T. Color Atlas and Manual of Microscopy for Criminalists, Chemists, and Conservators (Boca Raton-CRC Press, 2003).D’Agostino, A. et al. Environmental implications and evidence of natural products from dental calculi of a Neolithic–Chalcolithic community (central Italy). Sci. Rep. 11, 1–13 (2021).Article 

    Google Scholar 
    Frangiote-Pallone, S. & de Souza, L. A. Pappus and cypsela ontogeny in Asteraceae: structural considerations of the tribal category. Rev. Mex. Biodivers. 85, 62–77 (2014).Article 

    Google Scholar 
    Eglinton, G., Gonzalez, A. G., Hamilton, R. J. & Raphael, R. A. Hydrocarbon constituents of the wax coatings of plant leaves: a taxonomic survey. Phytochemistry 1, 89–102 (1962).Article 
    CAS 

    Google Scholar 
    Buckley, S. A., Stott, A. W. & Evershed, R. P. Studies of organic residues from ancient Egyptian mummies using high temperature-gas chromatography-mass spectrometry and sequential thermal desorption-gas chromatography-mass spectrometry and pyrolysis-gas chromatography-mass spectrometry. Analyst 124, 443–452 (1999).Article 
    CAS 

    Google Scholar 
    Hardy, K. et al. Neanderthal medics? Evidence for food, cooking, and medicinal plants entrapped in dental calculus. Naturwissenschaften 99, 617–626 (2012).Article 
    CAS 

    Google Scholar 
    Luong, S., Tocheri, M. W., Sutikna, T., Saptomo, E. W. & Roberts, R. G. Incorporating terpenes, monoterpenoids and alkanes into multiresidue organic biomarker analysis of archaeological stone artefacts from Liang Bua (Flores, Indonesia). J. Archaeol. Sci. Rep. 19, 189–199 (2018).
    Google Scholar 
    Luong, S. et al. Combined organic biomarker and use-wear analyses of stone artefacts from Liang Bua, Flores, Indonesia. Sci. Rep. 9, 1–17 (2019).Article 
    CAS 

    Google Scholar 
    Dabney, J., Meyer, M. & Pääbo, S. Ancient DNA damage. Cold Spring Harb. Perspect. Biol. 5, a012567 (2013).Article 

    Google Scholar 
    Mann, A. E. et al. Differential preservation of endogenous human and microbial DNA in dental calculus and dentin. Sci. Rep. 8, 1–15 (2018).Article 

    Google Scholar 
    Horrocks, M., Nieuwoudt, M. K., Kinaston, R., Buckley, H. & Bedford, S. Microfossil and Fourier Transform InfraRed analyses of Lapita and post-Lapita human dental calculus from Vanuatu, Southwest Pacific. J. R. Soc. N. Z. 44, 17–33 (2014).Article 

    Google Scholar 
    Radini, A., Nikita, E., Buckley, S., Copeland, L. & Hardy, K. Beyond food: the multiple pathways for inclusion of materials into ancient dental calculus. Am. J. Phys. Anthropol. 162, 71–83 (2017).Article 

    Google Scholar 
    Henry, A. G. Other microparticles: volcanic glass, minerals, insect remains, feathers, and other plant parts. In Handbook for the Analysis of Micro-Particles in Archaeological Samples 289–295 (Springer, Cham, 2020).MacKenzie, L., Speller, C. F., Holst, M., Keefe, K., & Radini, A. Dental calculus in the industrial age: human dental calculus in the Post-Medieval period, a case study from industrial Manchester. Quat. Int. https://doi.org/10.1016/j.quaint.2021.09.020 (2021).Radini, A., & Nikita, E. Beyond dirty teeth: Integrating dental calculus studies with osteoarchaeological parameters. Quat. Int. https://doi.org/10.1016/j.quaint.2022.03.003 (2022).Dobney, K. & Brothwell, D. A scanning electron microscope study of archaeological dental calculus. In Scanning Electron Microscopy in Archaeology BAR International Series (ed. & Olsen S), vol. 452, pp. 372–385 (Oxford, UK: BAR, 1988).Henry, A. G. & Piperno, D. R. Using plant microfossils from dental calculus to recover human diet: a case study from Tell al-Raqā’i, Syria. J. Archaeol. Sci. 35, 1943–1950 (2008).Article 

    Google Scholar 
    Wesolowski, V., de Souza, S. M. F. M., Reinhard, K. J. & Ceccantini, G. Evaluating microfossil content of dental calculus from Brazilian sambaquis. J. Archaeol. Sci. 37, 1326–1338 (2010).Article 

    Google Scholar 
    González-Guarda, E. et al. Multiproxy evidence for leaf-browsing and closed habitats in extinct proboscideans (Mammalia, Proboscidea) from Central Chile. Proc. Natl. Acad. Sci. USA 115, 9258–9263 (2018).Article 

    Google Scholar 
    Radley, J. A. Starch and its Derivatives (Chapman and Hall, London, 1968).Power, R. C., Salazar-García, D. C., Wittig, R. M. & Henry, A. G. Assessing use and suitability of scanning electron microscopy in the analysis of micro remains in dental calculus. J. Archaeol. Sci. 49, 160–169 (2014).Article 
    CAS 

    Google Scholar 
    Rottoli, M. & Castiglioni, E. Prehistory of plant growing and collecting in northern Italy, based on seed remains from the early Neolithic to the Chalcolithic (c. 5600–2100 cal BC). Veg. Hist. Archaeobot. 18, 91–103 (2009).Article 

    Google Scholar 
    Fiorentino, G. et al. Climate changes and human–environment interactions in the Apulia region of southeastern Italy during the Neolithic period. Holocene 23, 1297–1316 (2013).Article 

    Google Scholar 
    Rottoli, M., & Pessina, A. Neolithic agriculture in Italy: an update of archaeobotanical data with particular emphasis on northern settlements. In The Origins and Spread of Domestic Plants in Southwest Asia and Europe 157–170 (Routledge, 2016).Arobba, D., Panelli, C., Caramiello, R., Gabriele, M. & Maggi, R. Cereal remains, plant impressions and 14C direct dating from the Neolithic pottery of Arene Candide Cave (Finale Ligure, NW Italy). J. Archaeol. Sci. Rep. 12, 395–404 (2017).
    Google Scholar 
    Ucchesu, M., Sau, S. & Lugliè, C. Crop and wild plant exploitation in Italy during the Neolithic period: New data from Su Mulinu Mannu, Middle Neolithic site of Sardinia. J. Archaeol. Sci. Rep. 14, 1–11 (2017).
    Google Scholar 
    Scorrano, G. et al. Effect of Neolithic transition on an Italian community: Mora Cavorso (Jenne, Rome). Archaeol. Anthropol. Sci. 11, 1443–1459 (2019).Article 

    Google Scholar 
    De Angelis, F. et al. Exploring mobility in Italian Neolithic and Copper Age communities. Sci. Rep. 11, 1–14 (2021).Article 

    Google Scholar 
    Oxilia, G. et al. Exploring late Paleolithic and Mesolithic diet in the Eastern Alpine region of Italy through multiple proxies. Am. J. Phys. Anthropol. 174, 232–253 (2021).Article 

    Google Scholar 
    Fahmy, A. G. E. Palaeoethnobotanical studies of the Neolithic settlement in Hidden Valley, Farafra Oasis, Egypt. Veg. Hist. Archaeobot. 10, 235–246 (2001).Article 

    Google Scholar 
    Reed, K. From the field to the hearth: plant remains from Neolithic Croatia (ca. 6000–4000 cal bc). Veg. Hist. Archaeobot. 24, 601–619 (2015).Article 

    Google Scholar 
    Lucarini, G., Radini, A., Barton, H. & Barker, G. The exploitation of wild plants in Neolithic North Africa. Use-wear and residue analysis on non-knapped stone tools from the Haua Fteah cave, Cyrenaica, Libya. Quat. Int. 410, 77–92 (2016).Article 

    Google Scholar 
    García-Granero, J. J., Urem-Kotsou, D., Bogaard, A. & Kotsos, S. Cooking plant foods in the northern Aegean: microbotanical evidence from Neolithic Stavroupoli (Thessaloniki, Greece). Quat. Int. 496, 140–151 (2018).Article 

    Google Scholar 
    Bouby, L. et al. Early Neolithic (ca. 5850-4500 cal BC) agricultural diffusion in the Western Mediterranean: an update of archaeobotanical data in SW France. PLoS One 15, e0230731 (2020).Article 
    CAS 

    Google Scholar 
    Delhon, C., Binder, D., Verdin, P. & Mazuy, A. Phytoliths as a seasonality indicator? The example of the Neolithic site of Pendimoun, south-eastern France. Veg. Hist. Archaeobot. 29, 229–240 (2020).Article 

    Google Scholar 
    Lu, H. et al. Phytoliths analysis for the discrimination of foxtail millet (Setaria italica) and common millet (Panicum miliaceum). PLoS One 4, e4448 (2009).Article 

    Google Scholar 
    Celant, A. Indagini paleobotaniche su macroresti vegetali dai siti neo-eneolitici del territorio di Roma. In Roma prima del mito. Abitati e necropoli dal Neolitico alla prima età dei Metalli nel territorio di Roma (VI-III millennio a.C.) (eds Anzidei, A. P. & Carboni, C.) Vol. 2, 687–704 (Archaeopress Archaeol., 2020).Carra, M. et al. Plant foods in the Late Palaeolithic of Southern Italy and Sicily: Integrating carpological and dental calculus evidence. Quat. Int. https://doi.org/10.1016/j.quaint.2022.06.007 (2022) .Bednar, G. E. et al. Starch and fiber fractions in selected food and feed ingredients affect their small intestinal digestibility and fermentability and their large bowel fermentability in vitro in a canine model. J. Nutr. 131, 276–286 (2001).Article 
    CAS 

    Google Scholar 
    Hoover, R., Hughes, T., Chung, H. J. & Liu, Q. Composition, molecular structure, properties, and modification of pulse starches: a review. Food Res. Int. 43, 399–413 (2010).Article 
    CAS 

    Google Scholar 
    Wani, I. A. et al. Isolation, composition, and physicochemical properties of starch from legumes: a review. Starch‐Stärke 68, 834–845 (2016).Article 
    CAS 

    Google Scholar 
    Tayade, R., Kulkarni, K. P., Jo, H., Song, J. T. & Lee, J. D. Insight into the prospects for the improvement of seed starch in legume—a review. Front. Plant Sci. 10, 1213 (2019).Article 

    Google Scholar 
    Stafford, H. A. Distribution of tartaric acid in the leaves of certain angiosperms. Am. J. Bot. 46, 347–352 (1959).Article 
    CAS 

    Google Scholar 
    DeBolt, S., Cook, D. R. & Ford, C. M. L-Tartaric acid synthesis from vitamin C in higher plants. Proc. Natl. Acad. Sci. USA 103, 5608–5613 (2006).Article 
    CAS 

    Google Scholar 
    Fernández-García, E. et al. Carotenoids bioavailability from foods: from plant pigments to efficient biological activities. Food Res. Int. 46, 438–450 (2012).Article 

    Google Scholar 
    Gliszczyńska, A. & Brodelius, P. E. Sesquiterpene coumarins. Phytochem. Rev. 11, 77–96 (2012).Article 

    Google Scholar 
    Eerkens, J. The preservation and identification of Piñon resins by GC‐MS in pottery from the Western Great Basin. Archaeometry 44, 95–105 (2002).Article 
    CAS 

    Google Scholar 
    Barnard, H. et al. Mixed results of seven methods for organic residue analysis applied to one vessel with the residue of a known foodstuff. J. Archaeol. Sci. 34, 28–37 (2007).Article 

    Google Scholar 
    Wysocka, W., Przybył, A. & Brukwicki, T. The structure of angustifoline, an alkaloid of Lupinus angustifolius, in solution. Monatsh. Chem. 125, 1267–1272 (1994).Article 
    CAS 

    Google Scholar 
    Ohmiya, S., Saito, K., & Murakoshi, I. Lupine alkaloids. In The alkaloids: Chemistry and Pharmacology Vol. 47, 1–114) (Academic Press, 1995).Mancinotti, D., Frick, K. M. & Geu-Flores, F. Biosynthesis of quinolizidine alkaloids in lupins: mechanistic considerations and prospects for pathway elucidation. Nat. Prod. Rep. 39, 1423–1437 (2022).Article 
    CAS 

    Google Scholar 
    Silvestri, L., Achino, K. F., Gatta, M., Rolfo, M. F. & Salari, L. Grotta Mora Cavorso: physical, material and symbolic boundaries of life and death practices in a Neolithic cave of central Italy. Quat. Int. 539, 29–38 (2020).Article 

    Google Scholar 
    Steele, V. J., Stern, B. & Stott, A. W. Olive oil or lard?: distinguishing plant oils from animal fats in the archaeological record of the eastern Mediterranean using gas chromatography/combustion/isotope ratio mass spectrometry. Rapid Commun. Mass Spectrom. 24, 3478–3484 (2010).Article 
    CAS 

    Google Scholar 
    Buonasera, T. Investigating the presence of ancient absorbed organic residues in groundstone using GC–MS and other analytical techniques: a residue study of several prehistoric milling tools from central California. J. Archaeol. Sci. 34, 1379–1390 (2007).Article 

    Google Scholar 
    Luong, S. et al. Development and application of a comprehensive analytical workflow for the quantification of non-volatile low molecular weight lipids on archaeological stone tools. Anal. Met. 9, 4349–4362 (2017).Article 
    CAS 

    Google Scholar 
    Baeten, J., Jervis, B., De Vos, D. & Waelkens, M. Molecular evidence for the mixing of Meat, Fish and Vegetables in Anglo‐Saxon coarseware from Hamwic, UK. Archaeometry 55, 1150–1174 (2013).Article 
    CAS 

    Google Scholar 
    Evershed, R. P. Chemical composition of a bog body adipocere. Archaeometry 34, 253–265 (1992).Article 
    CAS 

    Google Scholar 
    Garnier, N., Bernal-Casasola, D., Driard, C. & Pinto, I. V. Looking for ancient fish products through invisible biomolecular residues in the roman production vats from the Atlantic. Coast J. Marit. Archaeol. 13, 285–328 (2018).Article 

    Google Scholar 
    Copley, M. S., Bland, H. A., Rose, P., Horton, M. & Evershed, R. P. Gas chromatographic, mass spectrometric and stable carbon isotopic investigations of organic residues of plant oils and animal fats employed as illuminants in archaeological lamps from Egypt. Analyst 130, 860–871 (2005).Article 
    CAS 

    Google Scholar 
    Reber, E. A. & Hart, J. P. Pine resins and pottery sealing: analysis of absorbed and visible pottery residues from central New York State. Archaeometry 50, 999–1017 (2008).Article 
    CAS 

    Google Scholar 
    Simopoulos, A. P. Omega‐3 fatty acids in wild plants, nuts and seeds. Asia Pac. J. Clin. Nutr. 11, S163–S173 (2002).Article 
    CAS 

    Google Scholar 
    Harris, W. S. et al. Stearidonic acid-enriched soybean oil increased the omega-3 index, an emerging cardiovascular risk marker. Lipids 43, 805–811 (2008).Article 
    CAS 

    Google Scholar 
    Gismondi, A., Rolfo, M. F., Leonardi, D., Rickards, O. & Canini, A. Identification of ancient Olea europaea L. and Cornus mas L. seeds by DNA barcoding. C. R. Biol. 335, 472–479 (2012).Article 
    CAS 

    Google Scholar 
    Steffens, W. & Wirth, M. Freshwater fish-an important source of n-3 polyunsaturated fatty acids: a review. Fish. Aquat. Sci. 13, 5–16 (2005).
    Google Scholar 
    Swanson, D., Block, R. & Mousa, S. A. Omega-3 fatty acids EPA and DHA: health benefits throughout life. Adv. Nutr. 3, 1–7 (2012).Article 
    CAS 

    Google Scholar 
    Wiermann, R., & Gubatz, S. Pollen wall and sporopollenin. In International Review of Cytology 35–72 (Academic Press, 1992).Cristiani, E., Radini, A., Edinborough, M. & Borić, D. Dental calculus reveals Mesolithic foragers in the Balkans consumed domesticated plant foods. Proc. Natl. Acad. Sci. USA 113, 10298–10303 (2016).Article 
    CAS 

    Google Scholar 
    Hardy, K. et al. Dental calculus reveals potential respiratory irritants and ingestion of essential plant-based nutrients at Lower Palaeolithic Qesem Cave Israel. Quat. Int. 398, 129–135 (2016).Article 

    Google Scholar 
    Radini, A. et al. Neanderthals, trees and dental calculus: new evidence from El Sidrón. Antiquity 90, 290–301 (2016).Article 

    Google Scholar 
    Lippi, M. M., Pisaneschi, L., Sarti, L., Lari, M. & Moggi-Cecchi, J. Insights into the Copper-Bronze Age diet in central Italy: plant microremains in dental calculus from Grotta dello Scoglietto (Southern Tuscany, Italy). J. Archaeol. Sci. Rep. 15, 30–39 (2017).
    Google Scholar 
    Modi, A. et al. Combined metagenomic and archaeobotanical analyses on human dental calculus: a cross-section of lifestyle conditions in a Copper Age population of central Italy. Quat. Int. https://doi.org/10.1016/j.quaint.2021.12.003 (2021).Warinner, C. et al. Pathogens and host immunity in the ancient human oral cavity. Nat. Genet. https://doi.org/10.1038/ng.2906 (2014).Lieverse, A. R. Diet and the aetiology of dental calculus. Int. J. Osteoarchaeol. 9, 219–232 (1999).Article 

    Google Scholar 
    Lukacs, J. R. & Largaespada, L. L. Explaining sex differences in dental caries prevalence: saliva, hormones, and “life‐history” etiologies. Am. J. Hum. Biol. 18, 540–555 (2006).Article 

    Google Scholar 
    Moore, P. D., Webb, J. A., & Collison, M. E. Pollen Analysis (Blackwell Scientific Publications, 1991).Borojević, K., Forenbaher, S., Kaiser, T. & Berna, F. Plant use at Grapčeva cave and in the eastern Adriatic Neolithic. J. Field Archaeol. 33, 279–303 (2008).Article 

    Google Scholar 
    Martin, L., Jacomet, S. & Tiebault, S. Plant economy during the Neolithic in a mountain context: the case of “Le Chenet des Pierres” in the French Alps (Bozel-Savoie, France). Veg. Hist. Archaeobot. 17, 113–122 (2008).Article 

    Google Scholar 
    Moser, D., Di Pasquale, G., Scarciglia, F. & Nelle, O. Holocene mountain forest changes in central Mediterranean: soil charcoal data from the Sila Massif (Calabria, southern Italy). Quat. Int. 457, 113–130 (2017).Article 

    Google Scholar 
    D’Agostino, A. et al. Pollen record of the Late Pleistocene–Holocene stratigraphic sequence and current plant biodiversity from Grotta Mora Cavorso (Simbruini Mountains, Central Italy). Ecol. Evol. 12, e9486 (2022).Radaeski, J. N., Bauermann, S. G. & Pereira, A. B. Poaceae pollen from Southern Brazil: distinguishing grasslands (campos) from forests by analyzing a diverse range of Poaceae species. Front. Plant Sci. 7, 1833 (2016).Article 

    Google Scholar 
    Turner, S. D. & Brown, A. G. Vitis pollen dispersal in and from organic vineyards: I. Pollen trap and soil pollen data. Rev. Palaeobot. Palynol. 129, 117–132 (2004).Article 

    Google Scholar 
    Marvelli, S., De’Siena, S., Rizzoli, E. & Marchesini, M. The origin of grapevine cultivation in Italy: the archaeobotanical evidence. Ann. Bot. 3, 155–163 (2013).
    Google Scholar 
    Riaz, S. et al. Genetic diversity analysis of cultivated and wild grapevine (Vitis vinifera L.) accessions around the Mediterranean basin and Central Asia. BMC Plant Biol. 18, 1–14 (2018).Article 

    Google Scholar 
    Arnold, C., Gillet, F., & Gobat, J. M. Situation de la vigne sauvage Vitis vinifera subsp. silvestris en Europe. Vitis 159–170 (1998).Terral, J. F. et al. Evolution and history of grapevine (Vitis vinifera) under domestication: new morphometric perspectives to understand seed domestication syndrome and reveal origins of ancient European cultivars. Ann. Bot. 105, 443–455 (2010).Article 

    Google Scholar 
    Buckley, S., Usai, D., Jakob, T., Radini, A. & Hardy, K. Dental calculus reveals unique insights into food items, cooking and plant processing in prehistoric central Sudan. PLoS One 9, e100808 (2014).Article 

    Google Scholar 
    Petrov, P. R., Popova, E. D. & Zlatanova, D. P. Niche partitioning among the red fox Vulpes vulpes (L.), stone marten Martes foina (Erxleben) and pine marten Martes martes (L.) in two mountains in Bulgaria. Acta Zool. Bulg. 68, 375–390 (2016).
    Google Scholar 
    Mikrjukov, K. A. Revision of genera and species composition of lower Centroheliozoa. II. Family Raphidiophryidae n. tam. Arch. Protistenkd. 147, 205–212 (1996).Article 

    Google Scholar 
    Cavalier-Smith, T. & von der Heyden, S. Molecular phylogeny, scale evolution and taxonomy of centrohelid heliozoa. Mol. Phylogen. Evol. 44, 1186–1203 (2007).Article 
    CAS 

    Google Scholar 
    Mertens, K. N., Rengefors, K., Moestrup, Ø. & Ellegaard, M. A review of recent freshwater dinoflagellate cysts: taxonomy, phylogeny, ecology and palaeocology. Phycologia 51, 612–619 (2012).Article 

    Google Scholar 
    Zlatogursky, V. V. Raphidiophrys heterophryoidea sp. nov. (Centrohelida: Raphidiophryidae), the first heliozoan species with a combination of siliceous and organic skeletal elements. Eur. J. Protist. 48, 9–16 (2012).Article 

    Google Scholar 
    Prokina, K. I. & Mylnikov, A. P. Centrohelid heliozoans from freshwater habitats of different types of South Patagonia and Tierra del Fuego, Chile. Inland Water Biol. 12, 10–20 (2019).Article 

    Google Scholar 
    Siemensma, F. J. & Roijackers, M. M. A study of new and little- known acanthocystid heliozoans, and a proposed division of the genus Acanthocystis (Actinopoda, Heliozoea). Arch. Protistenkd. 135, 197 (1988a).Article 

    Google Scholar 
    Siemensma, F. J. & Roijackers, M. M. The genus Raphidiophrys (Actinopoda, Heliozoea): scale morphology and species distinctions. Arch. Protistenkd. 136 237–248 (1988).Taylor, W.D. & Sanders, R. W. PROTOZOA. In Ecology and Classification of North American Freshwater Invertebrates (eds Thorp, J. H. & Covich, A. P.) 43–96 (Academic Press, 2001).Manconi, R., & Pronzato, R. Global diversity of sponges (Porifera: Spongillina) in freshwater. In Freshwater Animal Diversity Assessment 27–33 (Springer, Dordrecht, 2007).Malone, C. & Stoddart, S. The neolithic site of San Marco, Gubbio (Perugia), Umbria: survey and excavation 1985–7. Pap. Br. Sch. Rome 60, 1–69 (1992).Article 

    Google Scholar 
    Rottoli, M. La Marmotta, Anguillara Sabazia (RM). Scavi 1989. Analisi paletnobotaniche: prime risultanze, Appendice 1 M.A. In La Marmotta” (Anguillara Sabazia, RM). Scavi 1989. Un abitato perilacustre di età neolitica (eds. Fugazzola Delpino, M. A., D’Eugenio, G. & Pessina, A.) Bullettino di Paletnologia Italiana 84, 305–315 (1993).Pini, R. Late Neolithic vegetation history at the pile‐dwelling site of Palù di Livenza (northeastern Italy). J. Quat. Sci. 19, 769–781 (2004).Article 

    Google Scholar 
    Tinner, W. et al. Holocene environmental and climatic changes at Gorgo Basso, a coastal lake in southern Sicily, Italy. Quat. Sci. Rev. 28, 1498–1510 (2009).Article 

    Google Scholar 
    Bieniek, A. Archaeobotanical analysis of some early Neolithic settlements in the Kujawy region, central Poland, with potential plant gathering activities emphasized. Veg. Hist. Archaeobot. 11, 33–40 (2002).Article 

    Google Scholar 
    Tolar, T., Jacomet, S., Velušček, A. & Čufar, K. Plant economy at a late Neolithic lake dwelling site in Slovenia at the time of the Alpine Iceman. Veg. Hist. Archaeobot. 20, 207–222 (2011).Article 

    Google Scholar 
    D’Agostino, A. et al. Investigating plant micro-remains embedded in dental calculus of the Phoenician inhabitants of Motya (Sicily, Italy). Plants 9, 1395 (2020).Article 

    Google Scholar 
    Mercader, J. et al. Exaggerated expectations in ancient starch research and the need for new taphonomic and authenticity criteria. Facets 3, 777–798 (2018).Article 

    Google Scholar 
    Adojoh, O., Fabienne, M., Duller, R. & Osterloff, P. Taxonomy and phytoecology of palynomorphs and non-pollen palynomorphs: a refined compendium from the West Africa Margin. Biodivers. Int. J. 3, 188–200 (2019).Article 

    Google Scholar 
    Knapp, M., Clarke, A. C., Horsburgh, K. A. & Matisoo-Smith, E. A. Setting the stage building and working in an ancient DNA laboratory. Ann. Anat. 194, 3 (2012).Article 
    CAS 

    Google Scholar 
    Knapp, M., Lalueza-Fox, C. & Hofreiter, M. Re-inventing ancient human DNA. Investig. Genet. 6, 1 (2015).Article 

    Google Scholar 
    Gismondi, A. et al. Grapevine carpological remains revealed the existence of a Neolithic domesticated Vitis vinifera L. specimen containing ancient DNA partially preserved in modern ecotypes. J. Archaeol. Sci. 69, 75–84 (2016).Article 
    CAS 

    Google Scholar 
    Llamas, B. et al. From the field to the laboratory: controlling DNA contamination in human ancient DNA research in the high-throughput sequencing era. Sci. Technol. Archaeol. Res. 3, 1–14 (2017).Le Moyne, C. & Crowther, A. Effects of chemical pre-treatments on modified starch granules: recommendations for dental calculus decalcification for ancient starch research. J. Archaeol. Sci. Rep. 35, 102762 (2021).
    Google Scholar 
    Rolfo, M. F., Achino, K. F., Fusco, I., Salari, L. & Silvestri, L. Reassessing human occupation patterns in the inner central Apennines in prehistory: the case-study of Grotta Mora Cavorso. J. Archaeol. Sci. Rep. 7, 358–367 (2016).
    Google Scholar  More

  • in

    Biodiversity stabilizes plant communities through statistical-averaging effects rather than compensatory dynamics

    Empirical dataWe applied our theory to two datasets (Table 1): the plant survey dataset and the biodiversity-manipulated experiment dataset. The plant survey dataset contains nine sites of long-term grassland experiments across the United States (see also Hallett et al.10, and Zhao et al.23). Five of nine sites are from the Long Term Ecological Research (LTER) network (see Table 1). Plant abundances were measured either as biomass or as percent cover. In percent-cover cases, summed values can exceed 100% due to vertically overlapping canopies. All sites were sampled annually and were spatially replicated. We only used data of the plant survey dataset from unmanipulated control plots. Methods for data collection were constant over time.The biodiversity-manipulated experimental dataset comprises two long-term grassland experiments, BigBio and BioCON, at the Cedar Creek Ecosystem Science Reserve. Both experiments directly manipulated plant species number (1, 2, 4, 8, 16 for BigBio; and 1, 4, 9, 16 for BioCON). BioCON also contains different treatment levels for nitrogen and atmospheric CO2, but here only data from the ambient CO2 and ambient N treatments were used. We excluded plots with only one species. BigBio comprises 125 plots over 17 years, and BioCON comprises 59 plots over 22 years (Table 1).TheoryLet xi(t) denote the biomass of species i = 1, …, S at time t = 1, …, t and let μi = mean (xi (t)), σi = ({{mbox{sd}}})(xi (t)), and ({v}_{i}={sigma }_{i}^{2}) be the mean, standard deviation and variance of species i, computed through time. Let vij = cov (({x}_{i}left(tright),, {x}_{j}left(tright))) be the covariance, through time, of the dynamics of species i and j. Let xtot (left(tright)={sum }_{i}{x}_{i}(t)), ({mu }_{{{mbox{tot}}}}={sum }_{i}{mu }_{i}), ({v}_{{{mbox{tot}}}}={sum }_{i,j}{v}_{{ij}}), and ({{{{{{rm{sigma }}}}}}}_{{{{{{rm{tot}}}}}}}=sqrt{{v}_{{{{{{rm{tot}}}}}}}}). When population time series are uncorrelated, ({v}_{{{{{{rm{tot}}}}}}}={sum }_{i}{v}_{i}).As defined previously10,15, community stability is the inverse coefficient of variation of ({x}_{{{mbox{tot}}}}left(tright)), ({S}_{{{{{{rm{com}}}}}}}={mu }_{{{{{{rm{tot}}}}}}}/{sigma }_{{{{{{rm{tot}}}}}}}). Population stability is the inverse of weighted-average population variability9, ({sum }_{i}frac{{mu }_{i}}{{mu }_{{{{{{rm{tot}}}}}}}}{{CV}}_{i}={sum }_{i}frac{{mu }_{i}}{{mu }_{{{{{{rm{tot}}}}}}}}frac{{sigma }_{i}}{{mu }_{i}}={sum }_{i}frac{{sigma }_{i}}{{mu }_{{{{{{rm{tot}}}}}}}}), i.e, ({S}_{{pop}}={mu }_{{{{{{rm{tot}}}}}}}/{sum }_{i}{sigma }_{i}). The ratio of community stability over population stability is the Loreau-de Mazancourt asynchrony index14, Φ = ({sum }_{i}{sigma }_{i}/{sigma }_{{{{{{rm{tot}}}}}}}), so that$${S}_{{{{{{rm{com}}}}}}}=varPhi {S}_{{{{{{rm{pop}}}}}}}.$$
    (1)
    Now we suppose a hypothetical community with the same species-level variances and means as the original community but with species covariances equal to zero. Then, (1) becomes Scom_ip = (SAE)Spop, where ({S}_{{{{{{rm{com}}}}}}_{{{{{rm{ip}}}}}}}=frac{{mu }_{{{{{{rm{tot}}}}}}}}{sqrt{{sum }_{i}{v}_{i}}}=frac{{mu }_{{{{{{rm{tot}}}}}}}}{sqrt{{sum }_{i}{sigma }_{i}^{2}}}) is the value of community stability in the case of uncorrelated or independent populations and SAE is the component of Φ due to statistical averaging (here, “ip” stands for “independent populations”). The equation Scom_ip = (SAE)Spop can be interpreted as a definition of SAE. We then have$$SAE=frac{{S}_{{{{{{rm{com}}}}}}_{{{{{rm{ip}}}}}}}}{{S}_{{{{{{rm{pop}}}}}}}}=frac{{sum }_{i}{sigma }_{i}}{sqrt{{sum }_{i}{sigma }_{i}^{2}}}.$$
    (2)
    The compensatory effect is then the rest of Φ, i.e.,$$CPE=frac{{S}_{{{{{{rm{com}}}}}}}}{{S}_{{{{{{rm{pop}}}}}}}times SAE}=frac{{sum }_{i}{sigma }_{i}}{{sigma }_{{{{{{rm{tot}}}}}}}left({sum }_{i}{sigma }_{i}/sqrt{{sum }_{i}{sigma }_{i}^{2}}right)}=frac{sqrt{{sum }_{i}{sigma }_{i}^{2}}}{{sigma }_{{{{{{rm{tot}}}}}}}}.$$
    (3)
    Considering the classic variance ratio ({{{{{rm{varphi }}}}}}=frac{{V}_{{{{{{rm{tot}}}}}}}}{{sum }_{i}{V}_{i}}=frac{{sigma }_{{{{{{rm{tot}}}}}}}^{2}}{{sum }_{i}{sigma }_{i}^{2}}), our CPE is (1/sqrt{varphi }). Values CPE  > 1 (respectively, More

  • in

    Predicting potential global distribution and risk regions for potato cyst nematodes (Globodera rostochiensis and Globodera pallida)

    Evans, K., Franco, J. & De Scurrah, M. M. Distribution of species of potato cyst-nematodes in South America. Nematologica 21, 365–369. https://doi.org/10.1163/187529275×00103 (1975).Article 

    Google Scholar 
    Plantard, O. et al. Origin and genetic diversity of Western European populations of the potato cyst nematode (Globodera pallida) inferred from mitochondrial sequences and microsatellite loci. Mol. Ecol. 17, 2208–2218. https://doi.org/10.1111/j.1365-294X.2008.03718.x (2008).Article 
    CAS 

    Google Scholar 
    Price, J. A., Coyne, D., Blok, V. C. & Jones, J. T. Potato cyst nematodes Globodera rostochiensis and G. pallida. Mol. Plant Pathol. 22, 495–507. https://doi.org/10.1111/mpp.13047 (2021).Article 
    CAS 

    Google Scholar 
    CABI. Globodera rostochiensis (yellow potato cyst nematode). https://www.cabi.org/isc/datasheet/27034 (2021).CABI. Globodera pallida (white potato cyst nematode). https://www.cabi.org/isc/datasheet/27033 (2021).Ruthes, A. C. & Dahlin, P. The impact of management strategies on the development and status of potato cyst nematode populations in Switzerland: An overview from 1958 to present. Plant Dis. 106, 1096–1104. https://doi.org/10.1094/pdis-04-21-0800-sr (2021).Article 

    Google Scholar 
    Minnis, S. T. et al. Potato cyst nematodes in England and Wales—Occurrence and distribution. Ann. Appl. Biol. 140, 187–195. https://doi.org/10.1111/j.1744-7348.2002.tb00172.x (2002).Article 

    Google Scholar 
    Djebroune, A. et al. Integrative morphometric and molecular approach to update the impact and distribution of potato cyst nematodes Globodera rostochiensis and Globodera pallida (Tylenchida: Heteroderidae) in Algeria. Pathogens 10, 216. https://doi.org/10.3390/pathogens10020216 (2021).Article 

    Google Scholar 
    Vallejo, D. et al. Occurrence and molecular characterization of cyst nematode species (Globodera spp.) associated with potato crops in Colombia. PLoS One 16, e0241256. https://doi.org/10.1371/journal.pone.0241256 (2021).Article 
    CAS 

    Google Scholar 
    Hajjaji, A., Mhand, R. A., Rhallabi, N. & Mellouki, F. First report of morphological and molecular characterization of Moroccan populations of Globodera pallida. J. Nematol. 53, e2021-07. https://doi.org/10.21307/jofnem-2021-007 (2021).Article 
    CAS 

    Google Scholar 
    Camacho, M. J. et al. Potato cyst nematodes: Geographical distribution, phylogenetic relationships and integrated pest management outcomes in Portugal. Front. Plant Sci. 11, 9. https://doi.org/10.3389/fpls.2020.606178 (2020).Article 

    Google Scholar 
    Handayani, N. D. et al. Distribution, DNA barcoding and genetic diversity of potato cyst nematodes in Indonesia. Eur. J. Plant Pathol. 158, 363–380. https://doi.org/10.1007/s10658-020-02078-7 (2020).Article 
    CAS 

    Google Scholar 
    Bairwa, A. et al. Morphological and molecular characterization of potato cyst nematode populations from the Nilgiris. Indian J. Agric. Sci. 90, 273–278 (2020).Article 
    CAS 

    Google Scholar 
    Mburu, H. et al. Potato cyst nematodes: A new threat to potato production in East Africa. Front. Plant Sci. 11, 13. https://doi.org/10.3389/fpls.2020.00670 (2020).Article 

    Google Scholar 
    Altas, A., Evlice, E., Ozer, G., Dababat, A. & Imren, M. Identification, distribution and genetic diversity of Globodera rostochiensis (Wollenweber, 1923) Skarbilovich, 1959 (Tylenchida: Heteroderidae) populations in Turkey. Turk. Entomol. Derg. Turk. J. Entomol. 44, 385–397. https://doi.org/10.16970/entoted.740223 (2020).Article 

    Google Scholar 
    Dandurand, L.-M., Zasada, I. A., Wang, X. & Mimee, B. Current status of potato cyst nematodes in North America. Annu. Rev. Phytopathol. 57, 117–133. https://doi.org/10.1146/annurev-phyto-082718-100254 (2019).Article 
    CAS 

    Google Scholar 
    Blacket, M. J. et al. Molecular assessment of the introduction and spread of potato cyst nematode, Globodera rostochiensis, in Victoria, Australia. Phytopathology 109, 659–669. https://doi.org/10.1094/phyto-06-18-0206-r (2019).Article 
    CAS 

    Google Scholar 
    Sullivan, M. J., Inserra, R. N., Franco, J., Moreno-Leheude, I. & Greco, N. Potato cyst nematodes: Plant host status and their regulatory impact. Nematropica 37, 193–201 (2007).
    Google Scholar 
    Hodda, M. & Cook, D. C. Economic impact from unrestricted spread of potato cyst nematodes in Australia. Phytopathology 99, 1387–1393. https://doi.org/10.1094/phyto-99-12-1387 (2009).Article 
    CAS 

    Google Scholar 
    Koirala, S., Watson, P., McIntosh, C. S. & Dandurand, L. M. Economic impact of Globodera pallida on the Idaho economy. Am. J. Potato Res. 97, 214–220. https://doi.org/10.1007/s12230-020-09768-2 (2020).Article 

    Google Scholar 
    Trudgill, D. L., Elliott, M. J., Evans, K. & Phillips, M. S. The white potato cyst nematode (Globodera pallida)—A critical analysis of the threat in Britain. Ann. Appl. Biol. 143, 73–80. https://doi.org/10.1111/j.1744-7348.2003.tb00271.x (2003).Article 

    Google Scholar 
    Duan, Y. X. Plant Nematology (Science Press, 2011).
    Google Scholar 
    Peel, M. C., Finlayson, B. L. & McMahon, T. A. Updated world map of the Köppen–Geiger climate classification. Hydrol. Earth Syst. Sci. 11, 1633–1644. https://doi.org/10.5194/hess-11-1633-2007 (2007).Article 
    ADS 

    Google Scholar 
    Li, J. Suitable Risk Assessment for Six Potential Invasive Nematodes in China, Master thesis (Jilin Agriculture University, 2008).
    Google Scholar 
    Contina, J. B., Dandurand, L. M. & Knudsen, G. R. A spatiotemporal analysis and dispersal patterns of the potato cyst nematode Globodera pallida in Idaho. Phytopathology 110, 379–392. https://doi.org/10.1094/phyto-04-19-0113-r (2020).Article 
    CAS 

    Google Scholar 
    Guisan, A., Thuiller, W. & Zimmermann, N. E. Habitat Suitability and Distribution Models with Applications in R (Cambridge University Press, 2017). https://doi.org/10.1017/9781139028271.Book 

    Google Scholar 
    Phillips, S. J., Dudík, M. & Schapire, R. E. A maximum entropy approach to species distribution modeling. In Proceedings of the Twenty-First International Conference on Machine Learning 83. https://doi.org/10.1145/1015330.1015412 (Association for Computing Machinery, 2004).Phillips, S. J. & Dudík, M. Modeling of species distributions with Maxent: New extensions and a comprehensive evaluation. Ecography 31, 161–175. https://doi.org/10.1111/j.0906-7590.2008.5203.x (2008).Article 

    Google Scholar 
    Wan, J. et al. Predicting the potential geographic distribution of Bactrocera bryoniae and Bactrocera neohumeralis (Diptera: Tephritidae) in China using MaxEnt ecological niche modeling. J. Integr. Agric. 19, 2072–2082. https://doi.org/10.1016/S2095-3119(19)62840-6 (2020).Article 
    CAS 

    Google Scholar 
    Midmore, D. J. Potato production in the tropics. In The Potato Crop: The Scientific Basis for Improvement 728–793 (Springer Netherlands, 1992).Chapter 

    Google Scholar 
    Naika, S., de Jeude, J. V. L., de Goffau, M., Hilmi, M. & van Dam, B. Cultivation of Tomato: Production, Processing and Marketing (Digigrafi, 2005).
    Google Scholar 
    Chapman, M. A. Eggplant breeding and improvement for future climates. In Genomic Designing of Climate-Smart Vegetable Crops 257–276 (Springer, 2020).Chapter 

    Google Scholar 
    Management, D. o. C. List of National Agricultural Plant Quarantine Pests Distribution Administrative Areas. https://www.moa.gov.cn/nybgb/2019/201906/201907/t20190701_6320036.htm (Ministry of Agriculture and Rural Affairs of the People’s Republic of China, 2021).Elith, J. et al. A statistical explanation of MaxEnt for ecologists. Divers. Distrib. 17, 43–57. https://doi.org/10.1111/j.1472-4642.2010.00725.x (2011).Article 

    Google Scholar 
    Warren, D. L., Glor, R. E. & Turelli, M. ENMTools: A toolbox for comparative studies of environmental niche models. Ecography 33, 607–611. https://doi.org/10.1111/j.1600-0587.2009.06142.x (2010).Article 

    Google Scholar 
    Foot, M. A. The Ecology of Globodera pallida (Stone) Mulvey & Stone (Nematoda, Heteroderidae) at Pukekohe, New Zealand, Doctoral thesis (The University of Auckland, 1978).
    Google Scholar 
    Phillips, S. J., Dudík, M., & Schapire, R. E. Maxent software for modeling species niches and distributions (Version 3.4.1) [Internet]. http://biodiversityinformatics.amnh.org/open_source/maxent/. Accessed 24-10-2020.Merow, C., Smith, M. J. & Silander, J. A. Jr. A practical guide to MaxEnt for modeling species’ distributions: What it does, and why inputs and settings matter. Ecography 36, 1058–1069. https://doi.org/10.1111/j.1600-0587.2013.07872.x (2013).Article 

    Google Scholar 
    Wan, J., Wang, R., Ren, Y. & McKirdy, S. Potential distribution and the risks of Bactericera cockerelli and its associated plant pathogen Candidatus Liberibacter solanacearum for global potato production. Insects 11, 298. https://doi.org/10.3390/insects11050298 (2020).Article 

    Google Scholar 
    Cobos, M. E., Peterson, A. T., Barve, N. & Osorio-Olvera, L. kuenm: An R package for detailed development of ecological niche models using Maxent. PeerJ 7, e6281. https://doi.org/10.7717/peerj.6281 (2019).Article 

    Google Scholar 
    R Core Team. R: A language and environment for statistical computing. R Foundation for Statistical Computing. https://www.R-project.org/. Accessed 30-10-2020.Peterson, A. T., Papeş, M. & Soberón, J. Rethinking receiver operating characteristic analysis applications in ecological niche modeling. Ecol. Model. 213, 63–72. https://doi.org/10.1016/j.ecolmodel.2007.11.008 (2008).Article 

    Google Scholar 
    Warren, D. L. & Seifert, S. N. Ecological niche modeling in Maxent: The importance of model complexity and the performance of model selection criteria. Ecol. Appl. 21, 335–342. https://doi.org/10.1890/10-1171.1 (2011).Article 

    Google Scholar 
    ESRI. ArcGIS Desktop: Release 10 (Version 10.4.1) (Environmental Systems Research Institute). Accessed 24-10-2020.Kong, W., Li, X. & Zou, H. Optimizing MaxEnt model in the prediction of species distribution. J. Appl. Ecol. 30, 2116–2128. https://doi.org/10.13287/j.1001-9332.201906.029 (2019).Article 

    Google Scholar 
    Fischer, G. et al. Global Agro-ecological Zones (GAEZ v3.0). http://pure.iiasa.ac.at/id/eprint/13290/ (2012).Phillips, S. J., Anderson, R. P., Dudík, M., Schapire, R. E. & Blair, M. E. Opening the black box: An open-source release of Maxent. Ecography 40, 887–893. https://doi.org/10.1111/ecog.03049 (2017).Article 

    Google Scholar 
    da Silva, J. C. P., de Medeiros, F. H. V. & Campos, V. P. Building soil suppressiveness against plant-parasitic nematodes. Biocontrol Sci. Technol. 28, 423–445. https://doi.org/10.1080/09583157.2018.1460316 (2018).Article 

    Google Scholar 
    Kim, E., Seo, Y., Kim, Y. S., Park, Y. & Kim, Y. H. Effects of soil textures on infectivity of root-knot nematodes on carrot. Plant Pathol. J. 33, 66–74. https://doi.org/10.5423/ppj.Oa.07.2016.0155 (2017).Article 
    CAS 

    Google Scholar 
    Duyck, P.-F. et al. Niche partitioning based on soil type and climate at the landscape scale in a community of plant-feeding nematodes. Soil Biol. Biochem. 44, 49–55. https://doi.org/10.1016/j.soilbio.2011.09.014 (2012).Article 
    CAS 

    Google Scholar 
    Stanton, J. C., Pearson, R. G., Horning, N., Ersts, P. & ReşitAkçakaya, H. Combining static and dynamic variables in species distribution models under climate change. Methods Ecol. Evol. 3, 349–357. https://doi.org/10.1111/j.2041-210X.2011.00157.x (2012).Article 

    Google Scholar 
    Phillips, S. J., Anderson, R. P. & Schapire, R. E. Maximum entropy modeling of species geographic distributions. Ecol. Model. 190, 231–259. https://doi.org/10.1016/j.ecolmodel.2005.03.026 (2006).Article 

    Google Scholar 
    Burnham, K. P. & Anderson, D. R. Multimodel inference understanding AIC and BIC in model selection. Sociol. Methods Res. 33, 261–304. https://doi.org/10.1177/0049124104268644 (2004).Article 
    MathSciNet 

    Google Scholar 
    Din, A. U. et al. The impact of COVID-19 on the food supply chain and the role of e-commerce for food purchasing. Sustainability 14, 3074. https://doi.org/10.3390/su14053074 (2022).Article 
    CAS 

    Google Scholar 
    Lang, T. & McKee, M. The reinvasion of Ukraine threatens global food supplies. BMJ https://doi.org/10.1136/bmj.o676,o676,10.1136/bmj.o676 (2022).Article 

    Google Scholar 
    Hijmans, R. J. Global distribution of the potato crop. Am. J. Potato Res. 78, 403–412. https://doi.org/10.1007/bf02896371 (2001).Article 

    Google Scholar 
    Motti, R. The Solanaceae family: Botanical features and diversity. In The Wild Solanums Genomes 1–9 (Springer International Publishing, 2021).
    Google Scholar 
    Chytrý, M. et al. Projecting trends in plant invasions in Europe under different scenarios of future land-use change. Glob. Ecol. Biogeogr. 21, 75–87. https://doi.org/10.1111/j.1466-8238.2010.00573.x (2012).Article 

    Google Scholar 
    Lin, W., Cheng, X. & Xu, R. Impact of different economic factors on biological invasions on the global scale. PLoS One 6, e18797. https://doi.org/10.1371/journal.pone.0018797 (2011).Article 
    ADS 
    CAS 

    Google Scholar 
    Jones, L. M. et al. Climate change is predicted to alter the current pest status of Globodera pallida and G. rostochiensis in the United Kingdom. Glob. Change Biol. 23, 4497–4507. https://doi.org/10.1111/gcb.13676 (2017).Article 
    ADS 
    MathSciNet 

    Google Scholar 
    Kaczmarek, A., MacKenzie, K., Kettle, H. & Blok, V. C. Influence of soil temperature on Globodera rostochiensis and Globodera pallida. Phytopathol. Mediterr. 53, 396–405. https://doi.org/10.14601/Phytopathol_Mediterr-13512 (2014).Article 
    CAS 

    Google Scholar 
    Hearne, R., Lettice, E. P. & Jones, P. W. Interspecific and intraspecific competition in the potato cyst nematodes Globodera pallida and G. rostochiensis. Nematology 19, 463–475. https://doi.org/10.1163/15685411-00003061 (2017).Article 
    CAS 

    Google Scholar 
    Skelsey, P., Kettle, H., Mackenzie, K. & Blok, V. Potential impacts of climate change on the threat of potato cyst nematode species in Great Britain. Plant. Pathol. 67, 909–919. https://doi.org/10.1111/ppa.12807 (2018).Article 

    Google Scholar 
    Carlton, J. & Ruiz, G. M. Invasive Species: Vectors and Management Strategies (Island Press, 2003).
    Google Scholar 
    Singh, S. K., Paini, D. R., Ash, G. J. & Hodda, M. Prioritising plant-parasitic nematode species biosecurity risks using self organising maps. Biol. Invasions 16, 1515–1530. https://doi.org/10.1007/s10530-013-0588-7 (2014).Article 

    Google Scholar 
    Jiang, D., Chen, S., Hao, M. M., Fu, J. Y. & Ding, F. Y. Mapping the Potential global codling moth (Cydia pomonella L.) distribution based on a machine learning method. Sci. Rep. 8, 8. https://doi.org/10.1038/s41598-018-31478-3 (2018).Article 
    CAS 

    Google Scholar 
    Simberloff, D. et al. Impacts of biological invasions: What’s what and the way forward. Trends Ecol. Evol. 28, 58–66. https://doi.org/10.1016/j.tree.2012.07.013 (2013).Article 

    Google Scholar 
    Ravichandra, N. Nematodes of quarantine importance. In Horticultural Nematology 369–385 (Springer, 2014).
    Google Scholar  More

  • in

    Effects of diversity on thermal niche variation in bird communities under climate change

    Parmesan, C. Ecological and evolutionary responses to recent climate change. Annu. Rev. Ecol. Evol. Syst. 37, 637–669 (2006).Article 

    Google Scholar 
    Gilman, S. E., Urban, M. C., Tewksbury, J., Gilchrist, G. W. & Holt, R. D. A framework for community interactions under climate change. Trends Ecol. Evol. 25, 325–331 (2010).Article 

    Google Scholar 
    Devictor, V. et al. Differences in the climatic debts of birds and butterflies at a continental scale. Nat. Clim. Chang. 2, 121–124 (2012).Article 
    ADS 

    Google Scholar 
    Princé, K. & Zuckerberg, B. Climate change in our backyards: The reshuffling of North America’s winter bird communities. Glob. Change Biol. 21, 572–585 (2015).Article 
    ADS 

    Google Scholar 
    Brotons, L., Jiguet, F., Herando, S. & Lehikoinen, A. Bird communities and climate change. In Effects of Climate Change on Birds (eds Dunn, P. O. & Møller, A. P.) 221–235 (Oxford University Press, 2019).Chapter 

    Google Scholar 
    Chen, I. C., Hill, J. K., Ohlemüller, R., Roy, D. B. & Thomas, C. D. Rapid range shifts of species associated with high levels of climate warming. Science 333, 1024–1026 (2011).Article 
    ADS 
    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 
    Tylianakis, J. M., Didham, R. K., Bascompte, J. & Wardle, D. A. Global change and species interactions in terrestrial ecosystems. Ecol. Lett. 11, 1351–1363 (2008).Article 

    Google Scholar 
    Devictor, V., Julliard, R., Couvet, D. & Jiguet, F. Birds are tracking climate warming, but not fast enough. Proc. R. Soc. B Biol. Sci. 275, 2743–2748 (2008).Article 

    Google Scholar 
    Lehikoinen, A. et al. Wintering bird communities are tracking climate change faster than breeding communities. J. Anim. Ecol. 90, 1085–1095 (2021).Article 

    Google Scholar 
    McNaughton, S. J. Diversity and stability of ecological communities: A comment on the role of empiricism in ecology. Am. Nat. 111, 515–525 (1977).Article 

    Google Scholar 
    Loreau, M. et al. Biodiversity and ecosystem functioning: Current knowledge and future challenges. Science (80-. ) 294, 804–808 (2001).Article 
    ADS 
    CAS 

    Google Scholar 
    Loreau, M. & de Mazancourt, C. Biodiversity and ecosystem stability: A synthesis of underlying mechanisms. Ecol. Lett. 16, 106–115 (2013).Article 

    Google Scholar 
    Fonseca, C. R. & Ganade, G. Species functional redundancy, random extinctions and the stability of ecosystems. J. Ecol. 89, 118–125 (2001).Article 

    Google Scholar 
    Hodgson, D., McDonald, J. L. & Hosken, D. J. What do you mean, ‘resilient’?. Trends Ecol. Evol. 30, 503–506 (2015).Article 

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

    Google Scholar 
    Oksanen, J. et al. Community ecology package vegan, R package version 2.0-7 (2013).Laliberté, E., Legendre, P. & Shipley, B. FD: Measuring functional diversity from multiple traits, and other tools for functional ecology. R package (2014).García-Palacios, P., Gross, N., Gaitán, J. & Maestre, F. T. Climate mediates the biodiversity-ecosystem stability relationship globally. Proc. Natl. Acad. Sci. U. S. A. 115, 8400–8405 (2018).Article 
    ADS 

    Google Scholar 
    De Boeck, H. J. et al. Patterns and drivers of biodiversity-stability relationships under climate extremes. J. Ecol. 106, 890–902 (2018).Article 

    Google Scholar 
    Fridley, J. D. et al. The invasion paradox: Reconciling pattern and process in species invasions. Ecology 88, 3–17 (2007).Article 
    CAS 

    Google Scholar 
    Elton, C. S. The Ecology of Invasions by Plants and Animals (Methuen, 1958).Book 

    Google Scholar 
    Pigot, A. L., Trisos, C. H. & Tobias, J. A. Functional traits reveal the expansion and packing of ecological niche space underlying an elevational diversity gradient in passerine birds. Proc. R. Soc. B Biol. Sci. 283, 20152013 (2016).Article 

    Google Scholar 
    Pellissier, V., Barnagaud, J. Y., Kissling, W. D., Şekercioǧlu, Ç. H. & Svenning, J. C. Niche packing and expansion account for species richness–productivity relationships in global bird assemblages. Glob. Ecol. Biogeogr. 27, 604–615 (2018).Article 

    Google Scholar 
    Schipper, A. M. et al. Contrasting changes in the abundance and diversity of North American bird assemblages from 1971 to 2010. Glob. Change Biol. 22, 3948–3959 (2016).Article 
    ADS 

    Google Scholar 
    Jarzyna, M. A. & Jetz, W. A near half-century of temporal change in different facets of avian diversity. Glob. Change Biol. 23, 2999–3011 (2017).Article 
    ADS 

    Google Scholar 
    Catano, C. P., Fristoe, T. S., LaManna, J. A. & Myers, J. A. Local species diversity, β-diversity and climate influence the regional stability of bird biomass across North America. Proc. R. Soc. B Biol. Sci. 287, 20192520 (2020).Article 

    Google Scholar 
    Wang, S. et al. An invariability-area relationship sheds new light on the spatial scaling of ecological stability. Nat. Commun. 8, 1–8 (2017).ADS 

    Google Scholar 
    Pimm, S. L. & Redfearn, A. The variability of population densities. Nature 334, 613–614 (1988).Article 
    ADS 

    Google Scholar 
    Santangeli, A. & Lehikoinen, A. Are winter and breeding bird communities able to track rapid climate change? Lessons from the high North. Divers. Distrib. 23, 308–316 (2017).Article 

    Google Scholar 
    Sauer, J. R. et al. The first 50 years of the North American Breeding Bird Survey. Condor 119, 576–593 (2017).Article 

    Google Scholar 
    Meehan, T. D., Michel, N. L. & Rue, H. Spatial modeling of Audubon Christmas Bird Counts reveals fine-scale patterns and drivers of relative abundance trends. Ecosphere 10, e020707 (2019).Article 
    ADS 

    Google Scholar 
    Meller, K., Piha, M., Vähätalo, A. V. & Lehikoinen, A. A positive relationship between spring temperature and productivity in 20 songbird species in the boreal zone. Oecologia 186, 883–893 (2018).Article 
    ADS 

    Google Scholar 
    Lefcheck, J. S. & Duffy, J. E. Multitrophic functional diversity predicts ecosystem functioning in experimental assemblages of estuarine consumers. Ecology 96, 2973–2983 (2015).Article 

    Google Scholar 
    Alerstam, T. & Högstedt, G. Bird migration and reproduction in relation to habitats for survival and breeding. Scand. J. Ornithol. 13, 25–37 (1982).
    Google Scholar 
    Dingle, H. Migration: The Biology of Life on the Move (Oxford University Press, 1996).
    Google Scholar 
    Jones, P. D. et al. Hemispheric and large-scale land-surface air temperature variations: An extensive revision and an update to 2010. J. Geophys. Res. Atmos. https://doi.org/10.1029/2011JD017139 (2012).Article 
    ADS 

    Google Scholar 
    Rosenfeld, J. S. Functional redundancy in ecology and conservation. Oikos 98, 156–162 (2002).Article 

    Google Scholar 
    Bartley, T. J. et al. Food web rewiring in a changing world. Nat. Ecol. Evol. 3, 345–354 (2019).Article 

    Google Scholar 
    Thébault, E. & Loreau, M. Trophic interactions and the relationship between species diversity and ecosystem stability. Am. Nat. 166, E95–E114 (2005).Article 

    Google Scholar 
    Kokkoris, G. D., Jansen, V. A. A., Loreau, M. & Troumbis, A. Y. Variability in interaction strength and implications for biodiversity. J. Anim. Ecol. 71, 362–371 (2002).Article 

    Google Scholar 
    Vázquez, D. P. & Simberloff, D. Ecological specialization and susceptibility to disturbance: Conjectures and refutations. Am. Nat. 159, 606–623 (2002).Article 

    Google Scholar 
    Carvalheiro, L. G. et al. The potential for indirect effects between co-flowering plants via shared pollinators depends on resource abundance, accessibility and relatedness. Ecol. Lett. 17, 1389–1399 (2014).Article 

    Google Scholar 
    Morris, R. J., Lewis, O. T. & Godfray, H. C. J. Experimental evidence for apparent competition in a tropical forest food web. Nature 428, 310–313 (2004).Article 
    ADS 
    CAS 

    Google Scholar 
    Pace, M. L., Cole, J. J., Carpenter, S. R. & Kitchell, J. F. Trophic cascades revealed in diverse ecosystems. Trends Ecol. Evol. 14, 483–488 (1999).Article 
    CAS 

    Google Scholar 
    Dirzo, R. et al. Defaunation in the Anthropocene. Science (80-. ) 345, 401–406 (2014).Article 
    ADS 
    CAS 

    Google Scholar 
    Lindström, Å., Green, M., Paulson, G., Smith, H. G. & Devictor, V. Rapid changes in bird community composition at multiple temporal and spatial scales in response to recent climate change. Ecography (Cop.) 36, 313–322 (2013).Article 

    Google Scholar 
    Pearce-Higgins, J. W., Eglington, S. M., Martay, B. & Chamberlain, D. E. Drivers of climate change impacts on bird communities. J. Anim. Ecol. 84, 943–954 (2015).Article 

    Google Scholar 
    Socolar, J. B., Epanchin, P. N., Beissinger, S. R. & Tingley, M. W. Phenological shifts conserve thermal niches in North American birds and reshape expectations for climate-driven range shifts. Proc. Natl. Acad. Sci. 114, 12976–12981 (2017).Article 
    ADS 
    CAS 

    Google Scholar 
    Pollock, H. S., Brawn, J. D. & Cheviron, Z. A. Heat tolerances of temperate and tropical birds and their implications for susceptibility to climate warming. Funct. Ecol. https://doi.org/10.1111/1365-2435.13693 (2020).Article 

    Google Scholar 
    Wu, J. X., Wilsey, C. B., Taylor, L. & Schuurman, G. W. Projected avifaunal responses to climate change across the U.S. National Park System. PLoS ONE 13, 1–18 (2018).
    Google Scholar 
    Root, T. Environmental factors associated with avian distributional boundaries. J. Biogeogr. 15, 489 (1988).Article 

    Google Scholar 
    Zuckerberg, B. et al. Climatic constraints on wintering bird distributions are modified by urbanization and weather. J. Anim. Ecol. 80, 403–413 (2011).Article 

    Google Scholar 
    Newton, I. The Migration Ecology of Birds (Academic Press Inc., 2008).
    Google Scholar 
    La Sorte, F. A., Johnston, A. & Ault, T. R. Global trends in the frequency and duration of temperature extremes. Clim. Change 166, 1–14 (2021).Article 
    ADS 

    Google Scholar 
    Faurby, S. & Araújo, M. B. Anthropogenic range contractions bias species climate change forecasts. Nat. Clim. Change 8, 252–256 (2018).Article 
    ADS 

    Google Scholar 
    Jiguet, F., Brotons, L. & Devictor, V. Community responses to extreme climatic conditions. Curr. Zool. 57, 406–413 (2011).Article 

    Google Scholar 
    Tayleur, C. et al. Swedish birds are tracking temperature but not rainfall: Evidence from a decade of abundance changes. Glob. Ecol. Biogeogr. 24, 859–872 (2015).Article 

    Google Scholar 
    Clements, C. F. & Ozgul, A. Indicators of transitions in biological systems. Ecol. Lett. 21, 905–919 (2018).Article 

    Google Scholar 
    Scheffer, M. et al. Early-warning signals for critical transitions. Nature 461, 53–59 (2009).Article 
    ADS 
    CAS 

    Google Scholar 
    Fischer, J. et al. Functional richness and relative resilience of bird communities in regions with different land use intensities. Ecosystems 10, 964–974 (2007).Article 

    Google Scholar 
    Olivier, T., Thébault, E., Elias, M., Fontaine, B. & Fontaine, C. Urbanization and agricultural intensification destabilize animal communities differently than diversity loss. Nat. Commun. 11, 1–9 (2020).Article 
    ADS 

    Google Scholar 
    Michel, N. L. et al. Metrics for conservation success: Using the “Bird‐Friendliness Index” to evaluate grassland and aridland bird community resilience across the Northern Great Plains ecosystem. Divers. Distrib. 26, 1687–1702 (2020).National Audubon Society. The Christmas bird count historical results [online]. http://www.christmasbirdcount.org (2019).BirdLife International & NatureServe. Bird species distribution maps of the world. Version 4.0 (2015).Billerman, S. M., Keeney, B. K., Rodewald, P. G. & Schulenberg, T. S. Birds of the world (2020).BirdLife International. IUCN red list for birds. http://www.birdlife.org (2020).De Magalhães, J. P. & Costa, J. A database of vertebrate longevity records and their relation to other life-history traits. J. Evol. Biol. 22, 1770–1774 (2009).Article 

    Google Scholar 
    Wilman, H. et al. EltonTraits 1.0: Species-level foraging attributes of the world’s birds and mammals. Ecology 95, 2027–2027 (2014).Article 

    Google Scholar 
    IUCN. The IUCN red list of threatened species. Version 2019-2. http://www.iucnredlist.org (2019).Morelli, F., Benedetti, Y., Møller, A. P. & Fuller, R. A. Measuring avian specialization. Ecol. Evol. 9, 8378–8386 (2019).Article 

    Google Scholar 
    MacLean, S. A. & Beissinger, S. R. Species’ traits as predictors of range shifts under contemporary climate change: A review and meta-analysis. Glob. Change Biol. 23, 4094–4105 (2017).Article 
    ADS 

    Google Scholar 
    Jiguet, F., Gadot, A. S., Julliard, R., Newson, S. E. & Couvet, D. Climate envelope, life history traits and the resilience of birds facing global change. Glob. Change Biol. 13, 1672–1684 (2007).Article 
    ADS 

    Google Scholar 
    Julliard, R., Jiguet, F. & Couvet, D. Common birds facing global changes: What makes a species at risk?. Glob. Change Biol. 10, 148–154 (2004).Article 
    ADS 

    Google Scholar 
    Jetz, W., Thomas, G. H., Joy, J. B., Hartmann, K. & Mooers, A. O. The global diversity of birds in space and time. Nature 491, 444–448 (2012).Article 
    ADS 
    CAS 

    Google Scholar 
    Grimm, A. et al. Earlier breeding, lower success: Does the spatial scale of climatic conditions matter in a migratory passerine bird?. Ecol. Evol. 5, 5722–5734 (2015).Article 

    Google Scholar 
    Hijmans, R. J., Cameron, S. E., Parra, J. L., Jones, P. G. & Jarvis, A. Very high resolution interpolated climate surfaces for global land areas. Int. J. Climatol. 25, 1965–1978 (2005).Article 

    Google Scholar 
    Bates, D., Mächler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. J. Stat. Softw. https://doi.org/10.18637/jss.v067.i01 (2015).Article 

    Google Scholar 
    Villéger, S., Mason, N. W. H. & Mouillot, D. New multidimensional functional diversity indices for a multifaceted framework in functional ecology. Ecology 89, 2290–2301 (2008).Article 

    Google Scholar 
    Laliberté, E. & Legendre, P. A distance-based framework for measuring functional diversity from multiple traits. Ecology 91, 299–305 (2010).Article 

    Google Scholar 
    Barnagaud, J. Y. et al. Biogeographical, environmental and anthropogenic determinants of global patterns in bird taxonomic and trait turnover. Glob. Ecol. Biogeogr. 26, 1190–1200 (2017).Article 

    Google Scholar 
    Faith, D. P. Conservation evaluation and phylogenetic diversity. Biol. Conserv. 61, 1–10 (1992).Article 

    Google Scholar 
    Pinheiro, J., Bates, D., DebRoy, S., Sarkar, D. & R Core Team. nlme: Linear and nonlinear mixed effects model (2019).Barton, K. MuMIn: Multi-model inference. R package (2020).R Core Team. R: A language and environment for statistical computing. Version 3.5.3. http://www.r-project.org/ (2019).Devictor, V. et al. Functional biotic homogenization of bird communities in disturbed landscapes. Glob. Ecol. Biogeogr. 17, 252–261 (2008).Article 

    Google Scholar 
    Neutel, A. M., Heesterbeek, J. A. P. & De Ruiter, P. C. Stability in real food webs: weak links in long loops. Science (80-. ) 296, 1120–1123 (2002).Article 
    ADS 
    CAS 

    Google Scholar 
    Wallach, A. D. et al. Trophic cascades in 3D: Network analysis reveals how apex predators structure ecosystems. Methods Ecol. Evol. 8, 135–142 (2017).Article 

    Google Scholar  More

  • in

    The effects of temperature stress and population origin on the thermal sensitivity of Lymantria dispar L. (Lepidoptera: Erebidae) larvae

    In the autumn (November), L. dispar egg masses were collected at two sites: unpolluted and polluted forest. The first was a mixed oak forest at Kosmaj Mountain, 40 km south-east of Belgrade (coordinates 44°27′56″N 20°33′56″E). These woods are regarded as unpolluted because they are far from direct pollution and are part of the system of protected green areas around Belgrade, where the construction of industrial facilities and traffic infrastructure with potential negative effects on the environment is prohibited by legal regulations. The second site was Lipovica Forest (coordinates 44°38′11″N 20°24′12″E), with mixed Quercus frainetto and Quercus cerris trees, considered a polluted forest since it is located along the border of State Road 22, one of the most frequently used IB-class roads in Serbia.Collected egg masses were kept in a refrigerator at 4 °C until spring (March) when 200 eggs for each experimental group were set for hatching. After hatching in transparent Petri dishes (V = 200 mL), 10 first instar larvae were transferred and reared together at 23 °C with a 12:12 h light: dark photoperiod and relative humidity of 60%, until the third larval instar. Then, five 3rd instar larvae were reared together in the same Petri dish. After molting into the 4th instar, each larva was kept individually until the third day of the 5th instar, when they were sacrificed. Larvae were fed on an artificial diet designed for L. dispar42, and food was replaced every 48 h. Each experimental group contained between 50 and 60 larvae (Fig. 7).Figure 7A schematic figure of the experimental treatments.Full size imageThe optimal temperature for L. dispar larval development is 23 °C, and the control group was reared at this temperature. The highest summer temperature (2007–2010) measured in Serbian Quercus forests at a similar elevation was 28.4 °C, and the lowest 19.6 °C, while the average summer temperature was 26.3 °C43. Thus, we established variable temperature regimens that included brief (24 h) and daily (72 h) exposures to 28 °C. The control group of larvae were reared through the whole experiment on optimal 23 °C. Results of Huey et al.44 indicate that short term (daily) exposure to higher temperatures during development can increase both optimal temperature and maximal growth rate at the optimum, an example of beneficial thermal acclimation. In our previous research we found that induced thermotolerance modifies the activity of detoxifying enzymes in larvae originating from the polluted forest. We exposed L. dispar larvae in several experimental groups to that regime at 4th larval instar, with intention of analyze the effects of induce thermotolerance on observed parameters (ALP, ACP, hsp 70) in 5th instar larvae reared on optimal or elevated temperature28.At sacrifice on the third day of the 5th instar, the caterpillar midguts were dissected out on ice (n = 8–11 larval midguts per group for each enzyme assay). Midgut from single larvae was weighed and homogenized in insect physiological saline, as insect fluids have buffer values similar to vertebrates45. Homogenization was performed in ice-cold 0.15 M NaCl (final tissue concentration was 100 mg/mL in each sample), for 3 intervals of 10 s with a 15 s pause between them, at 5000 rpm, using Ultra Turrax homogenizer (IKA-Werke, Staufen, Germany). The homogenates were centrifuged for 10 min at 10,000 g at 4 ℃, and supernatants were used for enzyme assays and NATIVE gel electrophoresis. This protocol ensured that supernatants would contain cytosol and lysosomes.On the third day of the 5th instar, larval brain tissues were dissected out on ice and weighed. Pooled brain tissue (n = 30 brain tissues per experimental group) was diluted with 0.9% NaCl (1:9/w:V) and homogenized on ice at 5000 rpm during three 10 s intervals, separated by 15 s pauses (MHX/E Xenox homogenizer, Germany). Homogenates were centrifuged at 25,000 g for 10 min at 4 °C in an Eppendorf 5417R centrifuge (Germany). The supernatants were used for Western blotting and indirect non-competitive enzyme-linked immunosorbent assay (ELISA). Protein concentrations samples were determined using BSA as the standard46.A modified method by Nemec and Socha47 was used to determine the activity of ALP. The reaction mixture contained 0.1 M Tris HCl buffer pH 8.6, 5 mM MgCl2, midgut homogenate, and 5 mM p-nitrophenyl phosphate. During 30 min of incubation time at 30 ℃, the hydrolytic release of p-nitrophenol from p-nitrophenyl phosphate (pNPP) occurred under alkaline conditions.The reaction was stopped with 0.5 M NaOH, and the absorbance of p-nitrophenol was measured at 405 nm. Blank and non-catalytic probes were included. One unit of enzyme activity was defined as the amount of enzyme that released 1 mmol of p-nitrophenol per minute under the assay conditions.The same modified method of Nemec and Socha47 was employed to determine ACP activity, but under acidic conditions (0.1 M citrate buffer pH 5.6 was found optimal for L. dispar ACP), with a prolonged incubation time of 60 min. One unit of enzyme activity was defined as the amount of enzyme that released 1 μmol of p-nitrophenol per minute per mg of total protein. Total ACP activity determined in the midgut samples came from lysosomal ACP that ended up in the cytosol and non-lysosomal ACP, typically localized in the cytosol.Lysosomal ACP were detected indirectly48, under the same conditions, in a mixture containing the specific enzyme inhibitor NaF (50 mM). The absorbance determined at 405 nm is proportional to the activity of the non-lysosomal fraction of total ACP. The activity of the lysosomal fraction was obtained by subtracting not inhibited non-lysosomal acid phosphatases from the total phosphatase activity. Specific activities of ACP are given in mU per mg of total protein.A modified method by Allen et al.49 was used to detect ALP isoforms after native PAGE. Using 12% polyacrylamide gel, 10 μg protein aliquots per well were separated at 100 V and 4 ℃. The ALP isoform activity was visualized by soaking the gel in an incubation mixture consisting of 0.13% α-naphthyl phosphate, 100 mM Tris–HCl buffer (pH 8.6), and 0.1% Fast Blue B. The gels were incubated at room temperature until bands appeared.For ACP phosphatase detection, the same method of Allen et al.49 was also modified. After electrophoresis, the gel was washed with deionized water and equilibrated in 100 mM acetate buffer (pH 5.2) at 30 ℃. The nitrocellulose membrane was pre-soaked in 0.13% α-naphthyl phosphate dissolved in the same acetate buffer for 50 min at room temperature. The gel was covered with the membrane and incubated in a moist chamber for 60 min at 30 ℃. The membrane was soaked in 0.3% Fast Blue B stain dissolved in acetate buffer until bands became visible.Gels were scanned with a CanoScan LiDE 120 (Japan). The intensities of enzyme bands in the regions of ALP and ACP activities were analyzed using the ImageJ 1.42q software (U. S. National Institutes of Health, Bethesda, Maryland, USA).An indirect non-competitive ELISA was used to quantify the concentration of hsp70 in L. dispar brain tissue. Samples were diluted with carbonate-bicarbonate buffer (pH 9.6) and coated on a microplate (15 μg of tissue/well) (Multiwell immunoplate, NAXISORP, Thermo Scientific, Denmark) overnight at 4 °C, in the dark. The indirect non-competitive ELISA for L. dispar hsp70 was performed according to general practice: samples were first incubated with monoclonal anti-Hsp70 mouse IgG1 (dilution 1:5000) (clone BRM-22, Sigma Aldrich, USA) for 12 h at 4 °C, and then for 2 h at 25 °C with secondary anti-mouse IgG1 (gamma-chain)-HRP conjugate (dilution 1:5000) antibodies (Sigma Aldrich, USA). Chromogenic substrate 3, 3’, 5, 5’-Tetramethylbenzidine (TMB) was used as a visualizing reagent. Absorption was measured on a microplate reader (LKB 5060-006, Austria) at 450 nm. To enable statistically valid comparisons of experimental groups across multiple microplates, each microplate contained serial dilutions of standard hsp70 (recombinant hsp70, 50 ng/mL), used for the hsp70 standard curve, and homogenized brain tissues pulled by each treatment that were loaded on the microplates in a matched design, ensuring that each data point represented the mean of three replicates from each experimental group.Western blots were used to detect the presence of heat-shock protein 70 isoforms. Brain tissue homogenates were separated by SDS PAGE electrophoresis on 12% gels, according to Laemmli50. Protein transfer from the gel to the nitrocellulose membrane (Amersham Prothron, Premium 0.45 mm NC, GE Healthcare Life Sciences, UK) was left overnight at 40 V and 4 °C. Monoclonal anti-hsp70 mouse IgG1 (1:5000 dilution, clone BRM-22, Sigma Aldrich) and secondary mouse anti-mouse Hsp70 horseradish peroxidase conjugate antiserum (1:10,000 dilution, Sigma-Aldrich) were used for detection of hsp70 expression patterns in L. dispar larval brain tissue. Bands were visualized using chemiluminescence (ECL kit, Amersham).This study identified the hsp70 concentration in brain tissue and specific activities of total ACP and ALP in the larval midgut as the most promising biomarkers, which are sensitive and have consistent responses to thermal stress. These three biomarkers were combined into an IBR analysis according to Beliaeff and Burgeot51. The value of each biomarker (Xi) was standardized by the formula Yi = (Xi − mean)/SD, where Yi is the standardized biomarker response, and mean and SD were obtained from all values of the selected parameters. The next step was describing Zi as Zi = Yi or Zi = − Yi, depending on whether the temperature treatment caused induction or inhibition of the selected biomarkers. After finding the minimum value of Zi for each biomarker (min), the scores (Si) were computed as Si = Zi + |min|. Scores for biomarkers were used as the radius coordinates of the studied biomarker in the star plots. Star plot areas for the three-biomarker assembly, positioned in successive clockwise order—Hsp70, total ACP, and ALP, were obtained from the following formulas: ({A}_{i}=frac{{S}_{i}}{2*mathrm{sin}beta }left({S}_{i}*mathrm{cos}beta + {S}_{i+1}*mathrm{sin}beta right)), (beta = {mathrm{tan}}^{-1}left(frac{{S}_{i+1}*mathrm{sin}alpha }{{S}_{i}-{S}_{i+1}*mathrm{cos}alpha }right)),(alpha =2pi /n) radians (n is the number of biomarkers). The IBR values were calculated as follows:(IBR= sum_{i=1}^{n}{A}_{i}), where Ai is the area represented by two consecutive biomarkers on the star plot. Excel software (Microsoft, USA) was used to calculate IBR values and generate star plots.Statistical analyses were conducted in GraphPad Prism 6 (GraphPad Software, Inc., USA). Mean values ± standard errors of mean values (SEM) were calculated for the activity of enzymes, larval midgut mass, and the hsp70 concentration in brain tissue. D’Agostino-Pearson omnibus and Shapiro–Wilk tests were used to check the normality of data distribution. The effects of thermal treatments and their interaction on the variance of analyzed biomarkers in larvae from the polluted and the unpolluted forest were tested using two-way ANOVA with thermal treatments as fixed factors. For all comparisons, the level of significance was set at p  More

  • in

    Vulcanimicrobium alpinus gen. nov. sp. nov., the first cultivated representative of the candidate phylum “Eremiobacterota”, is a metabolically versatile aerobic anoxygenic phototroph

    Sampling, bacterial isolation, and colony screeningSamples were collected from thermally active sediments as evidenced by temperature and/or emission of steam in November 2010 and 2012 from Harry’s Dream (HD2 and HD3), Warren Cave (WC1, WC2, WC3, WC4, WC7, and WC8), Haggis Hole, (HH), Mammoth Cave (MC), Hut Cave (Hut), and Heroine Cave (HC) (fumarolic ice caves on Mt. Erebus volcano [21]). The samples were divided into four types: Mainly weathered basaltic/phonolitic sand (HD2, HD3, WC3, WC4, WC8 and Hut); pebbles and rock fragments (WC7, HH, and MC); black porous glassy materials that appeared to be solidified lava (WC1, WC2, and HC); and ash sediment (WC10) with little to no organic material (Table S1). The sediments were collected aseptically using sterile 50 mm conical tubes and immediately sealed. Additional information and environmental parameters on the sampling locations are provided in Table S1 and in Tebo et al. [21].To isolate bacteria found in these oligotrophic environments, we employed a culture strategy of long-term incubation in a nutrient-poor medium and screening of slow-growing colonies by direct PCR identification. Reasoner’s 2 A gellan gum medium (10% R2AG) [23] and FS1VG medium [24] were used for bacterial isolation. The 10% R2AG is a 10-fold diluted R2A broth (Nihon Seiyaku, Tokyo, Japan), solidified with 15 g/L gellan gum (Kanto chemical, Tokyo, Japan) with 2 g/L CaCl2. The 10% R2AG and FS1VG media were adjusted to a pH of 4.5 or 6.0, with or without 30 mg/L sodium azide. Samples used for isolation were selected from Warren Cave and Harry’s Dream, where the presence of relatively large numbers of bacteria ( >107/g) was confirmed in a previous report [21], and a total of nine samples, WD1, 2, 3, 4, 7, 8, and 10 and HD2 and 3, were used without any other pretreatment such as drying or dilution. For the sandy and ash samples with fine particles (WC3, 4, 8, 10, and HD2, 3), approximately 50 mg of sample were spread directly on plates. Glassy materials (WC1 and WC2) were embedded directly in plates using 2-3 pieces (approx. 200 mg). Pebbles and rock fragments (WC7) were crushed and approximately 50 mg of debris were spread directly onto the plates. All plates were incubated at 15 °C, 30 °C or 37 °C under dark conditions. New colonies were marked as they appeared and selection of the isolates was performed by picking only colonies that appeared after four weeks of incubation. We selected colonies around the sediment or in the cracks created during spreading using a magnifying glass.For identification, colonies were picked with a sterile toothpick, re-streaked, stabbed on fresh medium, and subsequently suspended in 20 μL sterilized 0.05 M NaOH. Suspensions were heated at 100 °C for 15 min, and supernatants were used as template DNAs for PCR. Partial 16 S rRNA gene sequences were amplified by PCR using commonly used bacterial primer set 27 F (5′-AGATTTGATCCTGGCTCAG-3′) and 1492 R (5′-GGTTACCTTGTTACGACTT-3′) or 536 R (5′-GTA TTA CCG CGG CTG CTG-3′) with TaKaRa ExTaq DNA polymerase (Takara Bio, Shiga, Japan) as previously described [23]. Sequencing was performed at Eurofins Genomics (Louisville, KY, USA), using a 3730xl DNA analyzer (Applied Biosystems, CA, USA). Sequence similarities with closest species were calculated using EZbiocloud’s Identify Service (https://www.ezbiocloud.net/identify). Subsequently, cells in the stab identified as “Ca. Eremiobacterota” by the direct PCR identification were serially diluted and stabbed onto new plates until multiple pure cultures of Eremiobacterota were obtained. The isolate was designated as strain WC8-2.Whole genome sequencing and annotationGenomic DNA was extracted from WC8-2 cells grown in 10% R2A broth (pH6.0) with air/CO2 (90:10, v/v) at 30 °C for 30 days under a 12/12 h light/dark regime with incandescent light (250 μmol m–2s–1), using a Puregene Yeast/Bact. Kit B (Qiagen, Germantown, MD, USA) [25]. Sequencing was performed by Macrogen Japan Corp., on a NovaSeq 6000 (Illumina, Inc., San Diego, CA, USA) and PacBio RSII (Pacific Biosciences of California, Inc. Menlo Park, CA, USA). The gap-free complete genome was assembled de novo using the Unicycler version 0.4.8 hybrid assembly pipeline with default settings [26]. Completeness and contamination levels were estimated using CheckM [27]. The genome was annotated using the DDBJ Fast Annotation and Submission Tool (DFAST) [28] and the BlastKOALA web server version 2.2, and was visualized using CGView Server [29] (http://cgview.ca/).Phylogenetic analysis of “Ca. Eremiobacterota”Identification of strain WC8-2 was performed using the Genome Taxonomy Database Toolkit (GTDB-Tk) (ver. 2.1.0), which produces standardized taxonomic labels that are based on those used in the Genome Taxonomy Database [30]. Terrabacterial genomes including “Ca. Eremiobacterota” MAGs and related genomes were retrieved from the Genome Taxonomy Database (GTDB) (July 2022) and the NCBI RefSeq database (July 2022). Full-length 16 S rRNA gene sequences were retrieved from the WC8-2 genome (WPS_r00030) and the NCBI database (Table S2). Multiple sequences were aligned using SINA (version 1.2.11) [31]. IQ-TREE version 1.6.12 [32] was used to build the phylogeny. ModelFinder [33] was used to determine the optimal evolutionary model for phylogeny building, which selected the TNe+I + G4 model. Branch support was calculated using 1000 ultrafast bootstraps [34]. The pairwise 16 S rRNA gene sequence similarities were determined using SDT software. Phylogenomic analysis based on 400 marker proteins was carried out using PhyloPhlAn v3.0 [35]. Diamond v5.2.32 [36], MAFFT v7.453 [37], and TrimAI were utilized for orthologs searching, multiple sequence alignment within the superphylum Terrabacteria, and gap-trimming, respectively. Gappy sites and sequences with >50% gaps were deleted from the alignments. IQ-TREE version 1.6.12 [32] was used to build the phylogenomic tree. ModelFinder [33] was used to determine the optimal evolutionary model for phylogeny building, which selected the LG + F + R9 model. These analyses were conducted using the “AOBA-B” super- computer (NEC, Tokyo, Japan) with 2CPUs (EPYC7702, AMD, CA, US) and 256GB of RAM. The related similarity of genomes between strain WC8-2 and relatives was estimated using average nucleotide identity (ANI) values, which were calculated using OrthoANI calculator in the EzBio-Cloud web service [38]. The related similarity between strain WC8-2 and its sister phyla with one representative from each class was assessed by pairwise Average Amino acid Identity (AAI) values using the online tools at the Kostas Konstantinidis Lab website Environmental Microbial Genomics Laboratory (http://enve-omics.ce.gatech.edu/aai/). The MAGs used in the tree are listed in Table S3.Phylogeny of photosynthesis- and “atmospheric chemosynthesis”- associated genesWe retrieved phototrophy- and “atmospheric chemosynthesis”-related protein sequences from the WC8-2 genome, “Ca. Eremiobacterota” MAGs, and known phototrophs genomes (Table S3) using the local BLAST server (SequenceServer 1.0.14 [39]) with reference sequences (Table S3) or annotated sequences as queries. Sequences were aligned using MUSCLE and poorly aligned regions were removed using Gblocks version 0.91b [40] or by manual inspection. The alignment sequences were concatenated into a single sequence. The ML tree was constructed using IQ-TREE with the best-hit evolutionary rate model: LG + I + G4 for HhyL, CbbL, BchXYZ and CbbL, LG + F + I + G4 for BchLNB, and LG + F + G4 for PufML, BchI, and BchD. All trees were visualized using iTOL (version 5.0) [41]. All sequences used in the trees are listed in Table S3.Analysis of bacterial communities in the fumarolic ice cavesEnvironmental DNA was extracted from the samples (0.1 g) (WC7, WC8, HD3, MC, Hut, HH, HC) using a DNeasy PowerSoil Kit (Qiagen, Valencia, CA, USA) following the manufacturer’s instructions. The V4 region of the 16 S rRNA gene was PCR amplified using primers with adapter sequences (V3-V4f_MIX) [42]. The PCR cycling was carried out using the following parameters: 94 °C for 2 min followed by 25 cycles at 94 °C for 30 s, 56 °C for 30 s, and 72 °C for 1 min with a final extension at 72 °C for 2 min. Library construction and sequencing were performed at the Bioengineering Lab (Kanagawa, Japan) using MiSeq (Illumina). Briefly, adaptor and primer regions were trimmed using the FASTX-Toolkit v0.0.13 (http://hannonlab.cshl.edu/fastx_toolkit). Read sequences of ≤40 bp with ambiguous bases and low-quality sequences (quality score, ≤Q20), together with their paired-end reads, were filtered out using Sickle v1.33 (https://github.com/najoshi/sickle). High-quality paired-end reads were merged using PEAR v0.9.10 with default settings [43]. Merged sequences of ≤245 and ≥260 bp were discarded using SeqKit v0.8.0 [44]. Operational taxonomic units (OTUs) were classified using QIIME v1.9.1 and the SILVA database (release 132) with 97% identity. To study the phylogeny of the OTUs assigned to the “Ca. Eremiobacterota”, a neighbor-joining (NJ) tree of the OTUs was constructed [45] as described above.Microscopic observationElectron microscopic observations of the cells were performed at Tokai Electron Microscopy (Nagoya, Japan). Briefly, the cells grown in 10% R2A broth (pH6.0) with air/CO2 (90:10, v/v) at 30 °C for 15 days under a 12/12 h light/dark regime with incandescent light (250 μmol m–2s–1) were fixed with 4% paraformaldehyde (PFA) and 4% glutaraldehyde (GA) in 0.1 M phosphate buffer (PB) at pH7.4, and postfixed with 2% OsO4 in 0.1 M PB. Cells were then dehydrated using graded ethanol solutions. The dehydrated cells were polymerized with resin, ultrathin sectioned, stained with 2% uranyl acetate, then secondary-stained with Lead stain solution. A transmission electron microscope (TEM) (JEM-1400Pus; JEOL, Tokyo, Japan) was used to observe the ultrathin sectioned cells at 100 kV acceleration voltage. To observe the negative-stained cells, PFA- and GA-fixed cells were adsorbed on formvar film-coated copper grids and stained with 2% phosphotungstic acid solution (pH 7.0) and observed using a TEM at 100 kV. DAPI (4,6-diamidino2-phenylindole) and Nile-Red staining was performed by incubating 0.1 mL cell suspension with a 1 mL staining solution (1 mg/L DAPI and 1 mg/L Nile Red in PBS buffer) for 10 min. The stained cells were observed under a fluorescence microscope (Olympus AX80T; Olympus Optical; Tokyo, Japan).Growth assayUnless otherwise noted, all cultures were grown in 100 mL butyl stopper- and screw-cap-sealed glass vials containing 50 mL liquid medium (pH6.0) at 30 °C. Growth was monitored by optical density at 600 nm (OD600) using a spectrophotometer (BioSpectrometer Basic; Eppendorf; Tokyo, Japan). Initial cell density was adjusted to 0.005 (OD600). Specific culture conditions are described below. All anaerobic growth tests were conducted with 100% N2 gas in the headspaces and supplemented with a reducing agent (0.3 g/L cysteine-HCl) and a redox indicator (1 mg/L resazurin).Optimal culture conditionsCell growth in different media was examined using 1, 10, 20, 100% (a full strength) R2A broth and Basal_YE with air/CO2 (90:10, v/v) under a 12/12 h light/dark regime with incandescent light (250 μmol m–2s–1) for 25 days. Basal_YE contained (l−1) 0.44 g KH2PO4, 0.1 g (NH4)2SO4, 0.1 g MgSO4.7H2O, 0.3 g yeast extract, and 1 ml trace element SL-8 [refer to DSMZ745]. Cell growth at different temperatures (10, 13, 20, 25, 30, 33 and 37 °C; pH6.0), pH (3.7, 4.5, 6.0, 7.0, 8.0, and 9.0), and NaCl (0, 1, 10, 20, and 30 g [l−1]; pH6.0) was examined using Basal_YE with air/CO2 (90:10, v/v) for 25 days under a 12/12 h light/dark regime with incandescent light (250 μmol m−2s−1). The following pH buffer solutions were used: acetic acid/sodium acetate for pH 4–6, K2HPO4/KH2PO4 for pH 6–8, sodium bicarbonate/sodium carbonate for pH 9–10. To examine colony formation on solid media, WC8-2 was streaked or stabbed on 10% R2AG medium (pH 6.0) and incubated at 30 °C for 30 days under a 100% air atmosphere.Optimal oxygen and carbon dioxide conditionsTo determine the preferred O2 concentration, WC8-2 was grown in Basal_YE (pH 6.0, 30 °C, no NaCl) in butyl stopper-sealed glass bottles with the atmosphere in the headspace adjusted to different N2/O2/CO2 ratios (70:20:10%, 80:10:10%, 89:1:10%, 90:0:10% v/v) after removing oxygen with 100% N2 gas. Cultures were incubated for 25 days under a 12/12 h light/dark regime with incandescent light (250 μmol m–2s–1). To determine the CO2 preference, the gas phase was adjusted to different N2/O2/CO2 ratios (70:20:10%,75:20:5%, 80:20:0% v/v) in sealed bottles or 100% air (plugged with a BIO-SILICO N-38 sponge plug; Shin-Etsu Polymer Co., Ltd, Tokyo, Japan; breathable culture-plug) under the same conditions as the O2 preference test.Photoorganoheterotrophic (or photoorganoautotrophic) and chemoorganoheterotrophic (or chemoorganoautotrophic) growthThe utilization of organic compounds as carbon sources/organic electron donors was tested in Basal medium (Basal_YE without yeast extract, pH 6.0) supplemented with one of the following sources (l-1): 0.3 ml of glycerol, or 0.3 g of sucrose, d-glucose, d-ribose, maltose, l-leucine, l-isoleucine, l-valine, l-serine, l-lysine, taurine, yeast extract or gellan gum, 1 ml of vitamin B12 solution (2 mg/L). Utilization was assessed by measuring growth of the cultures during a 25-day incubation at 30 °C in continuous light (250 μmol m–2s–1) for photoorganoheterotrophy, or continuous dark for chemoorganoheterotrophy under aerobic (air/CO2 [90:10, v/v]) and anaerobic (N2/CO2 [90:10, v/v]) conditions.Photolithoautotrophic and chemolithoautotrophic growthCells were inoculated into amended PSB2 [46] as described below or Basal medium with 5 mM Na2S or Na2S2O3 or 1% H2 (v/v; in the gas phase) as electron donors, and cultivated in continuous light (250 μmol m–2s–1) for photolithoautotrophic growth, or continuous dark for chemolithoautotrophic growth under aerobic (air/CO2 [90:10, v/v]) and anaerobic (N2/CO2 [90:10, v/v]) conditions for 60 days at 30 °C. The amended PBS2 contained (L-1): 0.5 g NH4Cl, 1.0 g KH2PO4, 0.2 g NaCl, 0.4 g MgSO4.7H2O, 0.05 g CaCl2.2H2O, 4.2 g NaHCO3, 1 ml trace element SL-8 [refer to DSMZ745], and 1 ml vitamin B12 solution (2 mg/L), pH6.0.Fermentative or anaerobic growthAnaerobic growth was examined in continuous dark in 20% R2A broth (pH 6.0) with N2/CO2 (90:10, v/v) supplemented with 5 mM Na2SO4, NaNO3, or dimethyl sulfoxide (DMSO) as electron acceptors for 60 days at 30 °C.Pigment assaysCells grown in Basal_YE (pH6.0) with air/CO2 (90:10, v/v) at 30 °C for 14 days (exponential growth phase) under continuous light (250 μmol m–2 s–1) and continuous dark were used for the pigment assays. The absorption spectrum was determined in a cell suspension in 60% (w/v) sucrose and in a 100% methanol extract using a spectrophotometer (V-630; JASCO, Tokyo, Japan) at 350–1100 nm. The BChl a concentration was determined spectroscopically in 100% methanol [47]. Dry cell weight was measured after harvested cells were washed twice with Milli-Q water and dried at 80 °C for 3 days. The extract was also analyzed by HPLC (NEXERA X2; Shimadzu; Kyoto, Japan) equipped with a 4.6 × 250 mm COSMOSIL 5C18-AR (Nakarai Taque; Tokyo, Japan) with isocratic elution of 92.5% (v/v) methanol in water at a flow rate of 1.0 mL/min. BChl a was monitored at 766 nm using a diode-array spectrophotometer detector (SPD-M20A; Shimadzu; Kyoto, Japan).Observation of taxisTo study phototaxis in WC8-2, cells were grown in 20% R2A broth with air/CO2 (90:10, v/v) for 14 days under a 12/12 h light/dark regimen. Cultures were transferred to tissue culture flasks (175 cm2, canted neck, Iwaki, Shizuoka, Japan). Light sources were ultraviolet (UV) at 395 nm (Linkman, Fukui, Japan), blue at 470 nm (CREE, Durham, NC, USA), green at 502 nm (Linkman), red at 653 nm (LENOO, Shinpei, Taiwan), and near-infrared (NIR) at 880 nm (LENOO). The light-emitting device was constructed by assembling LEDs on a breadboard with a power supply. Cultivations were illuminated with each wavelength using a light-emitting device from the underside and incubated at 20 °C for 18 h. Cells aggregating toward a light source was taken to indicate phototaxis. As a control, culture vessels were wrapped in aluminum foil to block light. Images and time-lapse video were captured using an iPhone 6 S camera.Stable carbon isotope ratio mass spectrometry (IRMS)The WC8-2 cells were grown in Basal_YE (pH6.0) with air/13CO2 [90:10, v/v] and air/unlabeled CO2 [90:10, v/v] under continuous light (250 μmol m–2s–1) for photoheterotroph and continuous dark for chemoheterotroph at 30 °C for 14 days (exponential growth phase). Approximately 10 mg culture biomass was collected, washed in HCl overnight, rinsed three times with deionized water, and placed into tin capsules. Stable carbon isotope ratios (δ13C) were analyzed at Shoko Science (Saitama, Japan) using a Delta V Advantage (EA-IRMS; Thermo Fisher Scientific, Bremen, Germany). The standard for C isotope ratio analysis was Vienna PeeDee Belemnite (VPDB). The δ13C values of 13CO2-cultivated cells exceeded the optimum calibration range of the instrument, but were used in this study to provide conclusive evidence that inorganic carbon was incorporated into the biomass.Quantitative reverse transcription PCR (qRT-PCR)Total RNA extraction and cDNA synthesisTotal RNA was extracted from cells grown in Basal_YE (pH6.0) with air/CO2 [90:10, v/v] in continuous light (250 μmol m–2s–1) for photoheterotrophic conditions and in continuous dark for chemoheterotrophic conditions at 30 °C for 14 days (exponential growth phase) using the Total RNA Purification Kit (Norgen, Biotek Corp, Ontario, Canada). DNA was removed from the extracted nucleic acids using an RNase-Free DNase I Kit (Norgen, Biotek Corp) according to the manufacturer’s protocol. The absence of DNA in the RNA samples was confirmed by PCR without reverse transcriptase. cDNA was generated from 500 ng total RNA using a TaKaRa PrimeScript™ 1st strand cDNA Synthesis Kit (TaKaRa Bio) with random hexamers according to the manufacturer’s protocol.Primer design, specificity and efficiencyThe following three photosynthesis- and CO2-fixation-related genes in the WC8-2 genome were selected for qRT-PCR: bchM encoding an enzyme involved in BChl synthesis, pufL encoding the anoxygenic Type II photochemical reaction centers L-subunit, and cbbL encoding the large subunit of type IE RuBisCO. The RNA polymerase subunit beta (rpoB) was used as a housekeeping reference gene. Primers for qRT-PCR were designed with Primer3 (v. 0.4.0) [48] with the following criteria: product size ranging from 80 to 150 bp, optimum Tm of 60 °C and GC content about 50 to 55%. Standard RT-PCR confirmed that each primer set amplified only a single product with expected size (data not shown), and the product was also sequenced using Sanger sequencing at Macrogen Japan Corp to confirm the candidate products. Primer efficiency was calculated for qRT-PCR using the slope of the calibration curve based on a 20-, 40-, 80-, 160-fold dilution series of cDNA samples [49]. In addition, the specificity of the primers was determined by the confirmation of a single peak in the melting curve. All information about the primers is shown in Table S4.qRT-PCRqRT-PCR was performed using a MiniOpticon Real-Time PCR System (Bio-Rad, Marnes la Coquette, France). The reaction mixture contained 10 μL TB Green Premix Ex Taq II (Tli RNaseH Plus, Takara Bio), 0.8 μL 10 mM primer, 2 μL of a 20-, 40-, 80-, 160-fold dilution series of cDNA, and 6.4 μL water. qRT-PCR was performed using the following protocol: denaturation at 95 °C for 30 s; denaturation and amplification at 95 °C for 5 s and 60 °C for 30 s, respectively (40 cycles). Fluorescence was measured at the end of the amplification step, and amplified products were examined by melting curve analysis from 60 to 95 °C. Each reaction was performed in three independent cultivations. Relative gene expression fold change was calculated using the comparative Ct method (2−ΔΔCt) [49]. Normalized expressions were used for reaction in dark. The 2−ΔΔCt values ≤0.5 were defined as downregulated and values ≥2.0 as upregulated, with P  More

  • in

    Reproductive performance and sex ratio adjustment of the wild boar (Sus scrofa) in South Korea

    Estes, J. A. Predators and ecosystem management. Wildl. Soc. Bull. 24, 390–396 (1996).
    Google Scholar 
    Licht, D. S., Millspaugh, J. J., Kunkel, K. E., Kochanny, C. O. & Peterson, R. O. Using small populations of wolves for ecosystem restoration and stewardship. Bioscience 60, 147–153 (2010).Article 

    Google Scholar 
    Schwartz, C. C., Swenson, J. E. & Miller, S. D. Large carnivores, moose, and humans: A changing paradigm of predator management in the 21st century. Alces J. Devot. Biol. Manag. Moose 39, 41–63 (2003).
    Google Scholar 
    Valente, A. M., Acevedo, P., Figueiredo, A. M., Fonseca, C. & Torres, R. T. Overabundant wild ungulate populations in Europe: Management with consideration of socio-ecological consequences. Mammal Rev. 50, 353–366 (2020).Article 

    Google Scholar 
    Lee, S. D. & Miller-Rushing, A. J. Degradation, urbanization, and restoration: A review of the challenges and future of conservation on the Korean Peninsula. Biol. Cons. 176, 262–276 (2014).Article 

    Google Scholar 
    Kodera, Y. Habitat selection of Japanese wild boar in Iwami district, Shimane Prefecture, western Japan. Wildl. Conserv. Jpn. 6, 119–129 (2001).
    Google Scholar 
    Ohashi, H. et al. Differences in the activity pattern of the wild boar Sus scrofa related to human disturbance. Eur. J. Wildl. Res. 59, 167–177 (2013).Article 

    Google Scholar 
    Ministry of Environment Republic of Korea. Management Plan of Pest Wild Boars. (Sejong, 2010).National Institute of Biological Resources. Analysis of Hunting Effect on Pest Animals and its Monitoring Techniques. (Incheon, 2017).Lee, S. M. & Lee, E. J. Diet of the wild boar (Sus scrofa): Implications for management in forest-agricultural and urban environments in South Korea. PeerJ 7, e7835 (2019).Article 

    Google Scholar 
    Bieber, C. & Ruf, T. Population dynamics in wild boar Sus scrofa: ecology, elasticity of growth rate and implications for the management of pulsed resource consumers. J. Appl. Ecol. 42, 1203–1213 (2005).Article 

    Google Scholar 
    Brogi, R. et al. Capital-income breeding in wild boar: A comparison between two sexes. Sci. Rep. 11, 1–10 (2021).Article 

    Google Scholar 
    Frauendorf, M., Gethöffer, F., Siebert, U. & Keuling, O. The influence of environmental and physiological factors on the litter size of wild boar (Sus scrofa) in an agriculture dominated area in Germany. Sci. Total Environ. 541, 877–882 (2016).Article 
    ADS 
    CAS 

    Google Scholar 
    Massei, G., Genov, P. V. & Staines, B. W. Diet, food availability and reproduction of wild boar in a Mediterranean coastal area. Acta Theriol. 41, 307–320 (1996).Article 

    Google Scholar 
    Sabrina, S., Jean-Michel, G., Carole, T., Serge, B. & Eric, B. Pulsed resources and climate-induced variation in the reproductive traits of wild boar under high hunting pressure. J. Anim. Ecol. 78, 1278–1290 (2009).Article 

    Google Scholar 
    Fonseca, C. et al. Reproduction in the wild boar (Sus scrofa Linnaeus, 1758) populations of Portugal. Galemys 16, 53–65 (2004).
    Google Scholar 
    Morreti, M. Birth distribution, structure and dynamics of a hunted mountain populatin of wild boars (Sus scrofa L.), Ticino, Switzerland. J. Mt. Ecol. 3, 192–196 (1995).
    Google Scholar 
    Rosell, C., Navas, F. & Romero, S. Reproduction of wild boar in a cropland and coastal wetland area: Implications for management. Anim. Biodivers. Conserv. 35, 209–217 (2012).Article 

    Google Scholar 
    Cellina, S. Effects of supplemental feeding on the body condition and reproductive state of wild boar Sus scrofa in Luxembourg PhD Dissertation thesis, University of Sussex, (2008).Gamelon, M. et al. High hunting pressure selects for earlier birth date: Wild boar as a case study. Evol. Int. J. Org. Evol. 65, 3100–3112 (2011).Article 

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

    Google Scholar 
    Fonseca, C., Da Silva, A., Alves, J., Vingada, J. & Soares, A. Reproductive performance of wild boar females in Portugal. Eur. J. Wildl. Res. 57, 363–371 (2011).Article 

    Google Scholar 
    Gaillard, J.-M., Brandt, S. & Jullien, J.-M. Body weight effect on reproduction of young wild boar (Sus scrofa) females: A comparative analysis. Folia Zool. (Brno) 42, 204–212 (1993).
    Google Scholar 
    Poteaux, C. et al. Socio-genetic structure and mating system of a wild boar population. J. Zool. 278, 116–125 (2009).Article 

    Google Scholar 
    Spitz, F., Valet, G. & Lehr Brisbin, I. Jr. Variation in body mass of wild boars from southern France. J. Mammal. 79, 251–259 (1998).Article 

    Google Scholar 
    Trivers, R. L. & Willard, D. E. Natural selection of parental ability to vary the sex ratio of offspring. Science 179, 90–92 (1973).Article 
    ADS 
    CAS 

    Google Scholar 
    Clutton-Brock, T. H., Albon, S. & Guinness, F. Parental investment in male and female offspring in polygynous mammals. Nature 289, 487–489 (1981).Article 
    ADS 

    Google Scholar 
    Hewison, A. M. & Gaillard, J.-M. Successful sons or advantaged daughters? The Trivers-Willard model and sex-biased maternal investment in ungulates. Trends Ecol. Evol. 14, 229–234 (1999).Article 
    CAS 

    Google Scholar 
    Clutton-Brock, T. H. & Iason, G. R. Sex ratio variation in mammals. Q. Rev. Biol. 61, 339–374 (1986).Article 
    CAS 

    Google Scholar 
    Fernández-Llario, P. & Mateos-Quesada, P. Body size and reproductive parameters in the wild boar Sus scrofa. Acta Theriol. 43, 439–444 (1998).Article 

    Google Scholar 
    Meikle, D. B., Drickamer, L. C., Vessey, S. H., Rosenthal, T. L. & Fitzgerald, K. S. Maternal dominance rank and secondary sex ratio in domestic swine. Anim. Behav. 46, 79–85 (1993).Article 

    Google Scholar 
    Servanty, S., Gaillard, J.-M., Allainé, D., Brandt, S. & Baubet, E. Litter size and fetal sex ratio adjustment in a highly polytocous species: The wild boar. Behav. Ecol. 18, 427–432 (2007).Article 

    Google Scholar 
    Mendl, M., Zanella, A. J., Broom, D. M. & Whittemore, C. T. Maternal social status and birth sex ratio in domestic pigs: An analysis of mechanisms. Anim. Behav. 50, 1361–1370 (1995).Article 

    Google Scholar 
    Cameron, E. Z. Facultative adjustment of mammalian sex ratios in support of the Trivers-Willard hypothesis: Evidence for a mechanism. Proc. R. Soc. Lond. Ser. B Biol. Sci. 271, 1723–1728 (2004).Article 

    Google Scholar 
    Clutton-Brock, T., Albon, S. & Guinness, F. Maternal dominance, breeding success and birth sex ratios in red deer. Nature 308, 358–360 (1984).Article 
    ADS 

    Google Scholar 
    Maillard, D. & Fournier, P. Timing and synchrony of births in the wild boar (Sus scrofa Linnaeus, 1758) in a Mediterranean habitat: The effect of food availability. Galemys 16, 67–74 (2004).
    Google Scholar 
    Bywater, K. A., Apollonio, M., Cappai, N. & Stephens, P. A. Litter size and latitude in a large mammal: the wild boar Sus scrofa. Mammal Rev. 40, 212–220 (2010).
    Google Scholar 
    Orłowska, L., Rembacz, W. & Florek, C. Carcass weight, condition and reproduction of wild boars harvested in north-western Poland. Pest Manag. Sci. 69, 367–370 (2013).Article 

    Google Scholar 
    Carranza, J. Sexual selection for male body mass and the evolution of litter size in mammals. Am. Nat. 148, 81–100 (1996).Article 

    Google Scholar 
    FernáNdez-Llario, P., Carranza, J. & Mateos-Quesada, P. Sex allocation in a polygynous mammal with large litters: The wild boar. Anim. Behav. 58, 1079–1084 (1999).Article 

    Google Scholar 
    McBride, G. The” teat order” and communication in young pigs. Animal Behaviour (1963).McBride, G., James, J. & Hodgens, N. Social behaviour of domestic animals. IV. Growing pigs. Anim. Sci. 6, 129–139 (1964).Article 

    Google Scholar 
    McBride, G., James, J. & Wyeth, G. Social behaviour of domestic animals VII. Variation in weaning weight in pigs. Anim. Sci. 7, 67–74 (1965).Article 

    Google Scholar 
    Geochang County. Geochang Statistical yearbook. (Geochang, 2015).Seoul Metropolitan Government. Seoul Statistical Yearbook. (Seoul, 2017).Animal and Plant Quarantine Agency and Ministry of Food and Drug Safety. Institutional Animal Care and Use Committee Standard Operation Guideline. (Gimcheon, 2020).Magnell, O. & Carter, R. The chronology of tooth development in wild boar – A guide to age determination of linear enamel hypoplasia in prehistoric and medieval pigs. Verterrinarija Ir Zootechnika. T. 40, 43–48 (2007).
    Google Scholar 
    Oroian, T. E., Oroian, R. G., Pasca, I., Oroian, E. & Covrig, I. Methods of age estimation by dentition in Sus scrofa ferus sp. Bull. UASVM Anim. Sci. Biotechnol. 67, 1–2 (2010).
    Google Scholar 
    Vericad, R. Fetal age, conception and birth period estimation on wild boar (Sus scrofa) in West Pyrenees. in Actas del XV Congresso Int. Fauna Cinegética y Silvestre. (Trujillo, 1983).Rosell, C., Fernández-Llario, P. & Herrero, J. The wild boar (Sus scrofa Linnaeus, 1758). Galemys 13, 1–25 (2001).
    Google Scholar 
    R core team. R: A language and environment for statistical computing v. 3.6.0 (Austria, 2019). More